You are on page 1of 7

Ind. Eng. Chem. Res.

1992,31, 581-587 58 1

Extraction and Fractionation of Fatty Acids from Oil Using an


Ultrafiltration Membrane
Jos T. F. Keurentjes,* Jos T. M. Sluijs, Robert-Jan H. Franssen, and Klaas van’t Riet
Food and Bioengineering Group, Department of Food Science, Wageningen Agricultural University,
P.O. Box 8129, 6700 EV Wageningen, The Netherlands

The removal of fatty acids from oil is an important step in edible oil refining. The classical caustic
refining process results in losses of triglycerides, which should preferably be avoided. In this study
a membrane-based liquid-liquid extraction is used in which these losses can be avoided by the use
of a selective extractant. A suitable extractant appears to be 1,2-butanediol. The distribution
coefficient for fatty acids varies with alkyl chain length from 0.7 for erucic acid (C22:l) up to 5.5
for caproic acid (C6). The fatty acids can be removed from the 1,Zbutanediol solution by the addition
of water. The system demixes, and the dispersion can then be separated into pure fatty acids and
a fatty acid free 1,2-butanediol/water mixture, to be reused as extractant after dewatering. Ex-
tractions have been performed in cellulose hollow fiber membrane modules. It appears that a major
part of the mass transfer resistance is situated inside the membrane wall. The overall mass transfer
coefficient varies from 7 X m/s for erucic acid u p to 5 X m/s for caproic acid. Using these
overall mass transfer coefficients, the required membrane surface area for a countercurrent extraction
can be calculated. An interesting feature is the fact that the mass transfer coefficients vary with
fatty acid chain length, which can be used for the fractionation of a fatty acid mixture. Two
mechanisms act in the same direction to obtain selectivity: an increasing mass transfer coefficient
and an increasing distribution coefficient with a decrease in fatty acid chain length.

Introduction membrane by one of the phases, cos 6 = 1. However, if


Liquid-liquid extraction is usually carried out by dis- this is not the case (cos 6 < l), the pressure which the
persing the extraction phase in the feed phase or vice versa. system can withstand is even lower.
In extraction towers, a large surface area to volume ratio For the separation of free fatty acids from oil, usually
can be obtained. After extraction, the dispersion is allowed sodium hydroxide is added to the oil in order to saponify
to coalesce and the two phases are separated. Coalescence, the fatty acids. Consequently, the soaps are separated
however, often causes severe problems in case of a very fine from the oil by high-speed separators (Applewhite, 1980;
dispersion, which occurs at intensive mixing to obtain a Torrey, 1983). The most important problem occurring in
large surface area per volume. Less intensive mixing re- this type of processes is the inclusion of triglycerides into
sults in a dispersion that can be separated more easily, but the soapstock. Usually, the amount of triglycerides in-
also results in a smaller surface area per volume, and hence cluded equals the amount of fatty acids (Braae, 1976), and
in a decreased extraction rate. Besides, backmixing or these triglycerides have to be considered as a loss. As an
flooding often limits the performance of extraction towers alternative, extractions have been proposed. However, the
(Lo and Baird, 1980; D’Elia et al., 1986). extractants used (mostly alcohols) show a low selectivity
Hollow fiber extractors offer the possibility to perform (Shah and Venkatesan, 1989; Pons and Eaves, 1967; Norris,
extractions without the formation of a dispersion (Prasad 1982; Uksila et al., 1966), which also causes losses of tri-
and Sirkar, 1988). Also, when the fibers are sufficiently glycerides.
small in diameter, a large extraction area per volume can In this study a series of extractants is tested. Factors
be obtained. Typical values are in the order of 103-105 affecting the extraction of fatty acids from an oil using
m2/m3(Kiani et al., 1984),which is higher than the values different homogeneous (symmetric) ultrafiltration mem-
found in packed towers (Lo and Baird, 1980). Flooding branes are investigated with a suitable extractant.
and backmixing do not occur in hollow fiber extractors. Theory
In most applications of hollow fiber extractors micro-
porous membranes are used (D’Elia et al., 1986; Prasad Overall Mass Transfer Coefficient. For a batch re-
and Sirkar,1988, Alexander and Callahan, 1987; Kim, 1984; circulation system, mass balance equations for the oil and
Callahan, 1988; Dekker et al., 1990). A static pressure is the extraction phase can be combined and linearized, re-
applied on the phase that does not wet the membrane. sulting in
However, in the case of a low interfacial tension between Ve + Vo/m
the two liquid phases, a rather unstable system is obtained
because the Laplace pressure in the pores is low. For
cylindrical pores this pressure is given by Here, Voand Ve are the volumes and C , and C, the con-
AP = (27 cos 6 ) / R (1) centrations of fatty acid in the oil and extraction phase,
respectively. The distribution coefficient m of the fatty
where AP is the pressure difference over the liquid-liquid acid over oil and the extraction phase is defined by m =
interface, y the interfacial tension between the two liquids, Ce,eq/Co,eq where the subscript eq indicates equilibrium.
6 the contact angle with the polymer surface, and R the A is the contact area between the two phases, and KOis
pore radius. Usually, there is complete wetting of the the overall mass transfer coefficient. When -In {(Ce,eq-
C e ) / C e , is plotted versus t , a straight line is obtained and
* To whom correspondence should be addressed at Akzo Re- KOcan%e calculated from the slope.
search, Dept. CRP, P.O.Box 9300, 6800 SB Arnhem, The Partial Mass Transfer Coefficients. For a hydro-
Netherlands. philic membrane with the extraction phase inside the
0888-5885/92/2631-0581$03.00/0 0 1992 American Chemical Society
582 Ind. Eng. Chem. Res., Vol. 31, No. 2,1992

membrane, the overall mass transfer can be written as Table I. Fatty Acids Used in This S t u d P
(Alexander and Callahan, 1987) estd chain
K;l = k;l + (mk,)-l + (mk,)-' (3) fatty acid manufacturer length, nm
caproic acid (C60) Aldrich 0.8
In this equation the overall mass transfer resistance (the capric acid (C100) Unichema 1.3
reciprocal of the overall mass transfer coefficient KJ is the myristic acid (C14:O) Aldrich 1.8
sum of the mass transfer resistance in the oil phase inside oleic acid (C181) Merck 2.5
the fibers (k,,-l),the mass transfer resistance in the ex- erucic acid (C22:l) Aldrich 3.1
traction phase outside the fibers ( ( m k e ) 3 ,and the re- OThe fatty acid chain length is calculated using a binding length
sistance of the extraction phase in the membrane wall of 0.154 nm for a C-C bond and a binding angle of logo.
((mk,)-').
Since the Reynolds number inside the fibers will be less Table 11. Membranes Used in Extraction Experiments
than 1, mass transfer inside the fibers is described by

a)
Wall
(Yang and Cussler, 1986) surface thickness, trade
membrane cutoff, D area, m2 1O-Sm name
113
-
kodf = 1.64( cellulose 6000 0.90 6.5 Cuprophan
(4) cellulose 6000 0.77 8 Cuprophan
DO cellulose 6000 0.70 11 Cuprophan
mod. cellulose 70000 0.177 16 Hemophan
in which is d fthe internal fiber diameter, 4 the fiber length, cellulose acetate 60 000 0.144 25 Diaph-an
Do the diffusion coefficient of the solute, and u, the average cellulose acetate 750000 0.062 85 PFlOO
oil flow velocity.
The resistance in the membrane wall (k,-l) is propor- phase. Once the overall mass transfer coefficient is lmown,
tional to the membrane thickness d , divided by the dif- it is possible to calculate the required surface area (A) for
fusivity De of the solute in the liquid filling the pores. With a given A and x i / x o .
a correction for the effective diffusion distance by the
tortuosity 7 and a correction for the polymer volume by Materials
the porosity e, the resistance in the membrane is given by Fatty acids with different chain lengths were used.
(Dahuron and Cussler, 1988) Some relevant characteristics are summarized in Table I.
km-l = drnT/Dee (5) All fatty acids used were reagent grade, except oleic acid,
which was technical grade.
The mass transfer outside the fibers can be calculated Extractions were carried out in hollow fiber membrane
using equations for mass transfer outside fibers or tubes devices containing different membrane materials, provided
as given in literature. All those equations are of the form by ENKA Membrana AG, FRG. The membranes used are
given in Table 11.
All experiments were carried out at 30 "C. The water
used was demineralized, and all other reagents were ana-
lytical grade and purchased from Merck (FRG). The oil
where u, is the extractant flow velocity, d the outer fiber phase was soybean oil of edible quality (Rhenus Inc., The
diameter and v the viscosity of the extraction liquid. The Netherlands).
constantsa and b differ with the system investigated (Yang
and Cussler, 1986; Dahuron and Cussler, 1988, Prasad and Experimental Section
Sirkar, 1987; Kreith, 1973), resulting in k, values that may Distribution coefficients were determined as follows.
vary by a factor of 100. For this reason the best way for After intensive mixing of 100 mL of fatty acid containing
the determination of the k, values is to measure KO,to soybean oil with 100 mL of extractant, the fatty acid
determine k, and k, and calculate k, therefrom, using eqs concentration was determined in both phases and mass
3-5. balances were set up in order to check the recovery. Fatty
Extraction. For a continuous extraction, the amount acid concentrations were determined by titration of a 1-g
of fatty acids that can be transferred from the oil phase sample in 20 mL of ethanol with a 0.5 N sodium hydroxide
into the extraction phase is determined by both the dis- solution. Water concentrations were measured using an
tribution coefficient m and the ratio of the oil phase and automatic Karl Fischer titration apparatus (Mettler).
extraction phase flow (F, and F,, respectively), as combined Membrane extractions were carried out using hollow
in the stripping factor A: fiber membrane modules. The experimental setup is de-
A = mFe/Fo (7) picted in Figure l. The oil phase is circulated inside and
the extraction phase outside the fibers and the polar ex-
When we consider the extraction to take place by a traction phase wets the fiber wall. In order to avoid
countercurrent process (which can be relatively easily emulsion formation by extractant permeating into the oil
achieved in a hollow fiber membrane device), the number phase, the oil phase was circulated at a static pressure of
of theoretical transfer units (NTU) equals the height of 0.5 X 105N m-2. Before use the membranes are rinsed with
the extractor (H)divided by the height of a transfer unit demineralized water.
(HTU) (Beek and Muttzall, 1975). This can be written Fatty acid concentrations in fractionation experiments
as were determined on a Carlo Erba gas chromatograph with
a 5-m CP-Si1 5 CB capillary column (Chrompack, The
A Netherlands) with a cold on-column injection system using
A-1 F O
hexadecane as an internal standard.
(NTU = H/HTU) (8) Results and Discussion
with xi/x, as the ratio between the inflow and the outflow Distribution Coefficients. For several extractanta the
concentration in the oil phase and p the density of the oil distribution coefficient for oleic acid over the extractant
Ind. Eng. Chem. Res., Vol. 31, No. 2, 1992 583
m 1-1
1.2, 1

0.8 k
. . 4 ' o00 2 4 6 8 1 0 1 2

% H20 in 1,2-butanediol

Figure 2. Effect of water content of 1,2-butanediol on the distri-


bution coefficient of oleic acid over soybean oil and 1,2-butanediol.
in 1-1

61 9
I \
Figure 1. Experimental setup for membrane extraction experi-
menta. 4i
4
\ \
Table 111. Distribution Coefficients of Oleic Acid over
Soybean Oil and Different Extractants"
miscibility
extractant m with oil
0 6 10 14 18 22
formamide 0.04 -- number o f C-atoms of btty acid
glycerol 0.16 --
1,3-butanediol 0.29 -- Figure 3. Effect of fatty acid alkyl chain length on the distribution
2,3-butanediol ++ coefficient of fatty acids over 1,2-butanediol (containing 2% water)
and soybean oil.
1,I-butanediol 0.12 --
1,2-butanediol 1.08 -- addition results in demixing of the system, yielding a
2-butene-1,4-diol 0.08 --
methanol 1.10 +- dispersion with fatty acids as the dispersed phase and
methanol/acetonitrile/water 75/20/5 (v/v/v) 0.45 -+ l,Zbutanediol/water as the continuous phase. The
methanol/furfural/water 15/20/5 (v/v/v) ++ amount of water required for demixing increases up to
"Miscibility of the extractant with oil is indicated by --, not 55% for capric acid, and is almost independent of the fatty
miscible; + -, partially miscible, and + +, completely miscible. acid content. Below the water content requried for de-
mixing, fatty acids are completely miscible with the 1,2-
and soybean oil is determined at an oleic acid concentra- butanediol mixture. After phase separation a fatty acid
tion of 20% (v/v). The results are given in Table 111. phase and a l,Zbutanediol/water phase are obtained. The
From Table I11 it appears that several extractants found fatty acid phase then contains 2% 1,2-butanediol,whereas
in the literature, like methanol and mixtures of methanol the 1,2-butanediol/water mixture contains less than 0.05%
with water and acetonitrile or furfural, are partially or fatty acids. These values are virtually independent of the
completely miscible with oil. For an effective extraction, nature of the fatty acid, indicating that a higher mutual
however, it is necessary that the mutual solubility is very solubility in the case of shorter fatty acids is compensated
low. This requirement, together with the highest distri- by an increased water content required for demixing. After
bution coefficient, can be met using 1,2-butanediol. dewatering this stream can be reused as extraction liquid
The distribution coefficient appears to be almost inde- (Keurentjes et al., 1991).
pendent of the fatty acid concentration and temperature, Mass Transfer Coefficients. Because of the low in-
but decreases with an increase in water content (Figure terfacial tension between fatty acid containing oil and
2). The chain length of the fatty acid strongly influences 1,Bbutanediol (which vanishes to zero a t a fatty acid
the distribution coefficient (Figure 3), varying from 0.7 for content of 35% (v/v) in the oil phase), and probably also
erucic acid up to 5.5 for caproic acid. It could be expected because of incomplete wetting by the water phase, it ap-
that short-chain fatty acids are more soluble in butanediol peared impossible to use the PFlOO cellulose acetate
than the long-chain fatty acids, due to the smaller ratio membrane for these extractions. Even at a low static
between hydrophobic tail and hydrophilic head group. pressure (0.05 X lo5 N/m2) the oil phase permeates
The solubility of soybean oil in 1,2-butanediolappeared through the membrane. Applying the same static pressure
to be below the threshold of detection using refractive on the water circuit results in permeation of the water
indices (i.e., <0.1% w/w). The solubility of 1,2-butanediol phase. The other membranes listed in Table I1 could be
in soybean oil was measured to be 1% (w/w) in the case used for the extraction experiments.
where butanediol is used without water. Since cellulose For the determination of the overall mass transfer re-
membranes are not allowed to be used with dry solvenb, sistance, experiments have been carried out with different
it is necessary to add some water to the butanediol. When fatty acids in soybean oil and the Cuprophan cellulose
1,Bbutanediol contains 2% (w/w) water, the solubility of membranes. Measuring the concentration of fatty acid in
1,2-butanediol in soybean oil decreases to a value below the extraction circuit with time yields Figure 4a as a typical
the threshold of detection using refractive indices (<0.1% result. Linearizingthis curve according to eq 2 then yields
w/w). Figure 4b, from which the overall mass transfer coefficient
When 1,2-butanediol is used as extractant, extracted can be calculated. The overall mass transfer coefficient
oleic acid can be removed from the l,&butanediol phase increases with a decrease in fatty acid chain length (Figure
by the addition of water up to 35% (w/w). This water 5).
684 Ind. Eng. Chem. Res., Vol. 31, No. 2,1992

-"

n 00

40 ~ 0 0

0O 04 7 20-

04 I I , , , , I
0 400 800 1xK) oi I I , I I , , I 1 . 8 .

time ( m i n l
a Ln ( I - A)
cem
I.IO-~ mI
0.321

OZ4]
0164

I
o d . 1 ' I ( 1
0 40 80 120
t i m e Lminl
Figure 4. (a, top) Fatty acid concentration versus time in a typical
extraction experiment of oleic acid from oil. (b, bottom) Linearized
concentration versus time for determination of K, Q representa the membrane thickness 1
0'm I
volume correction term of eq 2.
Figure 8. l/Kn versus dry membrane thickness for the three Cu-
K,( l i i a m / s ) prophan membranes.

D , ~i 10-l1m2/sI
32

&O-1 \ \

fatty acid Length Inm)


0 02 04 06 08
Figure 5. Overall mass transfer coefficient versus fatty acid chain
length (8-pm Cuprophan membrane). LEA [-]
rpWe

K: (1o6s/m) Figure 9. Diffusion inside the membrane matrix using the different
membranes of Table II, except for the PFlOO membrane. Values are
compared to values obtained with diffusion models proposed in the
literature.
therefore be expected that the intercept of Figure 6 rep-
resents the maas transfer resistance inside the membrane
wall only. Also, a plot of l/Ko versus the membrane wall
thicknets (Figure 8) gives a straight line through the origin,
indicating that indeed almost all resistance against mass
0
transfer is situated inside the membrane wall. This ap-
0 2 4 6 8 10 peared to be the case for all fatty acids used, except for
vpm IIsm13'P caproic acid (C6:0),where the resistance inside the fiber
Figure 6. l/Kn versus for determination of the partial mass wall only represents 60% of the overall mass transfer re-
sistance. According to eq 3, the mass transfer resistances
-
transfer resistance inside the fibers (8-pm Cuprophan membrane).
inside the fiber wall and at the extractant side of the
In order to determine the contribution of the partial membrane can be neglected when m m.
mass transfer resistance inside the fibers to the overall Diffusion in the Membrane Matrix. From mea-
mass transfer resistance, 1/K, is plotted versus u;lI3 ac- surements of the different partial mass transfer resistances,
cording to eq 4 (Figure 6). Only a slight dependence is the diffusion coefficients inside the membrane matrix can
found, and extrapolation to an infinitely large velocity still be calculated. These are given in Figure 9. The diffusion
yields a significant mass transfer resistance (the intercept coefficientsinside the cellulose membrane matrix obtained
represents the sum of the resistances inside the fiber wall for different fatty acids can be compared to the coefficients
and in the extraction phase). When 1/K, is plotted versus as obtained from different models that have been proposed
-l (Figure 7), no dependence is found, indicating that no in the literature to relate the diffusion coefficient in so-
&nificant resistance is situated outside the fibers. It can lution (calculated according to the Wilke-Chang rela-
Ind. Eng. Chem. Res., Vol. 31, No. 2, 1992 585

ai\\ I

_/---

- 2

0 1 ' 3
" 5
" 7" 9' ' 1 11
A I-]
0 2 4 6 8 10
[lo3mz.s/kg 1
FR
Figure 10. Membrane surface area required for the separation of Figure 11. Selectivity between different fatty acid combinations
oleic acid from soybean oil using the Diaphan membrane as a func- versus surface area to feed ratio at a stripping factor of 2 for oleic
tion of the stripping factor at different required separation effcien- acid using the Diaphan membrane.
cies.
-CCo
tionship (Wilke and Chang, 1955) to the diffusion coeffi-
cient inside a homogeneous membrane matrix. The results
of these calculations are also included in Figure 9. Most
models use the porosity, tortuosity, and pore diameter of
the membrane material (0.65 (Peppas and Reinhart,1983),
1.9 (Sakai et al., 1987), and 3.4 nm (Klein et al., 1979), -
respectively), except the model proposed by Peppas and
Reinhart (Peppas and Reinhart, 1983), which requires
[-le,
+i
information about the molecular structure of the polymer.
It appears that most models predict the diffusion
coefficient within a factor of 4. The curve following from 0 2 4 6 E 10
the theory of Faxen (Faxen, 1923) is differently shaped and time I1000 S I
predicts too steep a decrease of the diffusion coefficient Figure 12. Fractionation of a mixture of C6,C10,and C181 (k65)
with chain length, whereas both the models of Faxen and in a batch extraction.
Satterfield (Satterfield et al., 1973) differ more than a
factor of 4 for the longest fatty acids. From these results membranes, mass transfer coefficients are small as com-
it can be concluded that all models give predictions in the pared to transport in solution. This implies that relatively
range of the experimental values for the diffusivity inside large membrane surface areas are needed for an extraction.
the membrane matrix. However, none of them results in However, it yields another opportunity, which is the ability
an exact fit of the experimental data, which may be partly to achieve a fractionation of a fatty acid mixture, since
due to deviations in the estimates for the fatty acid and different fatty acids exhibit different mass transfer coef-
pore diameter. It seems that the models can, at best, make ficients, as is clearly shown in Figure 5. For this frac-
a reasonable first guess of the diffusion coefficient. tionation the selectivity (a)with respect to two components
Whenever values are needed that are more reliable than A and B can be defined as follows (Sandell, 1968):
that, they will have to be determined experimentally.
Countercurrent Extraction. Once the overall mass
transfer coefficients are obtained, it is possible to calculate
the required membrane surface area for a given extraction.
As an example, the surface area required for the separation The ratios xi,A/~04 and xia/xoa in this expression can be
of oleic acid from oil using the Diaphan membrane is calculated according to eq 8. In Figure 11calculated se-
calculated for different extraction efficiencies (xi/xJ versus lectivities between three different fatty acids (CS,C10, and
the stripping factor A. The overall mass transfer coeffi- C181) are given as a function of A/Fo at a stripping factor
cient for this separation is the measured one, being 8 X A = 2 for C181. It should be noted that the stripping
m/s. The results of these calculations are shown in factor increases with a decrease in fatty acid chain length
Figure 10. It appears that an increase in the stripping as a result of the increased distribution coefficient. This
factor has a strong influence on the required surface area implies that for fractionation two mechanisms act at the
up to values of about 3. A further increase does not result same time in the same direction: a difference in retarda-
in a further decrease in the required surface area. tion by the membrane and a difference in solubility in the
These data reveal that for a full-scale extraction (90% extraction liquid. Figure 11shows that an increase of the
fatty acid removal from 5000 kg/h oil) about 40OOO m2 of permeated amount (which increases with an increase in
membrane area are required. The costa for the extraction A/FJ leads to decreased selectivities. Obviously, at 100%
of copper from oil using the same cellulose hollow fiber permeation no separation is achieved.
membranes aa used in this study for a 10000-m2system By way of example, we give, in Figure 12, the removal
to treat 5OOO kg/h have been estimated to be between 0.01 of a low concentration C6 from a fatty acid mixture con-
and 0.03 $/kg (Keurentjes et al., 1990). The increase of sisting of C10 and C18:l dissolved in soybean oil. In this
the costs of a membrane plant are estimated to be pro- experiment a batch extraction with 0.77-m2membrane area
portional to Ao.6(Kloosterman et al., 1987). These figures is used. The volumes of the oil and the 1,2-butanediol
indicate that the costs for this fatty acid extraction will phase were 4.9 X lo4 and 4.3 X m3, respectively. It
be between 0.025 and 0.07 $/kg. appears that in a relatively short contact time (1h) the
Fractionation. Because of the low interfacial tension caproic acid concentration is decreased by 80%, whereas
between oil and 1,2-butanediol,it is necessary to use rather the oleic acid and capric acid concentration are only de-
"dense" membranes, i.e., with small pores. For such creased by 22% and 40%, respectively. The lines in Figure
586 Ind. Eng. Chem. Res., Vol. 31, No. 2, 1992

(21 P2 of water, resulting in demixing of the solution into a dis-


persion containing fatty acids as the dispersed and 1,2-
butanediol/water as the continuous phase. The fatty acid
F1,
phase then contains 2% 1,2-butanediol, and the 1,2-bu-
R2
tanediol contains less than 0.05% fatty acids.
Extractions have been performed in hollow fiber mem-
brane modules containing cellulose and cellulose acetate
membranes. Due to the low interfacial tension between
oil and 1,2-butanediol, it is necessary to use membranes
Fe3 (31 P3 with small pores. From mass transfer measurements it
Figure 13. Three-stage membrane system for fractionation of a follows that a major part of the mass transfer resistance
fatty acid mixture dissolved in an organic solvent. is situated inside the membrane matrix. Overall mass
transfer coefficients vary significantly for different fatty
Table IV. Concentrations in the Different Streams As acids, varying from 7 x IO* m/s for C22:l up to 5 X lo-’
Indicated in Figure 12 Starting with a 1:l:l Ratio in the m/s for C6.
Feed; A/F,= 4300 mz s/kg, A(C18:l) = 2 Using the mass transfer data it is possible to calculate
concentration, % the required membrane surface area for a given extraction.
stage C6 C10 C18:l As a result of the high mass transfer resistance inside the
1 permeate (Pl) 60 34 6 membrane wall, this membrane surface area is large.
retentate (Rl) 8 33 59 Fractionation of fatty acid mixtures can be achieved on
2 permeate (P2) 25 57 18 the basis of differences in overall mass transfer coefficients
retentate (R2) 2 21 77 and differences in solubility in 1,2-butanediol. These two
3 permeate (P3) 70 22 8 effects act in the same direction: short-chain fatty acids
retentate (R3) 25 57 18
have higher mass transfer coefficients and are better
soluble in 1,2-butanediol than the long-chain fatty acids.
12 are calculated according to eq 2, using mass transfer
coefficients from Figure 5. The agreement between ex- Acknowledgment
perimental and predicted values indicates that the indi-
vidually measured mass transfer Coefficients are not in- We thank the Dutch Program Committee for Industrial
fluenced significantlyby the presence of other fatty acids. Biotechnology (PCIB) for their financial support and B.
In this example only one extraction stage is used. More Hasenack for technical assistance.
stages will obviously result in almost pure products. Most Nomenclature
fatty acid fractionation procedures are based on crystal-
lization (e.g., panning and pressing (Zilch, 1979) and the A = membrane surface area, m2
hydrophilization process (Stein, 1968)),which require re- a = constant
action times of several days (Zilch, 1979). The major ad- b = constant
vantage of the present membrane fractionation procedure C = fatty acid concentration, kg/m3
D = fatty acid diffusion coefficient, m2/s
is found in the time required for separation. Neither of d = external fiber diameter, m
the crystallization processes results in pure products and df = internal fiber diameter, m
requires several stages, as does this membrane fractiona- d, = membrane thickness, m
tion, thus stressing the advantage of short reaction times. F = flow, kg/s
Using the data of Figure 11, a three-membrane coun- H = height of the extractor, m
tercurrent extraction system (Figure 13) can be used to HTU = height of a transfer unit, m
fractionate a 1:l:l mixture of C6, C10, and C18:l to pro- KO= overall mass transfer coefficient, m/s
duce one stream almost free of C6 (R2), one stream almost lz = partial mass transfer coefficient, m/s
free of C181 (P3), and one intermediate stream enriched lf = fiber length, m
in C10 (P2 and R3 together). The results are given in m = distribution coefficient
Table IV. NTU = number of theoretical transfer units
However, the two-stage extraction is more complicated AP = Laplace pressure, Pa
than shown in Figure 13. The fatty acids of the 1,2-bu- R = pore radius, m
tanediol phase have to be reextracted to another oil phase, t = time, s
since it is not possible to extract fatty acids from 1,2-bu- V = volume, m3
tanediol to fatty acid free 1,2-butanediol because of a u = velocity, m/s
transmembrane flow due to osmotic effects. Reextraction x = fatty acid concentration in oil phase, kg/m3
cy = selectivity
can easily be achieved by the addition of water to the
butanediol phase as described above, This operation yields y = interfacial tension, N/m
t = porosity
a fatty acid free butanediol phase and pure fatty acids A = stripping factor
which then can be dissolved in any organic solvent, thus T = tortuosity
serving as the feed of stage 3. 6 = contact angle, deg
v = viscosity, m2/s
Conclusions p = density, kg/m3

From this work it can be concluded that it is possible Subscripts


to extract fatty acids from an oil stream when a suitable o = oil phase, outflow oil phase
extractant can be found. It appeared that 1,a-butanediol e = extraction phase
can be used for this purpose. The distribution coefficient eq = equilibrium state
varies from 0.7 for erucic acid up to 5.5 for caproic acid. 1 = inflow oil phase
The mutual solubility of oil and 1,Zbutanediol is less than m = membrane
0.1% in the presence of 2% of water in the system. Re- A = component A
covery of the fatty acids can be achieved by the addition B = component B
Ind. Eng. Chem. Res. 1992,31, 587-592 587
Registry No. Caproic acid, 142-62-1; capric acid, 334-48-5; Kreith, F. Principles of Heat Transfer;Harper and Row New York,
myristic acid, 544-63-8; oleic acid, 112-80-1; erucic acid, 112-86-7; 1973.
1,3-butanediol, 584-03-2. Lo, T. C.; Baird, M. H. I. Liquid-liquid extraction. In Kirk-Othmer
Encyclopedia of Chemical Technology;Grayson, M., Ed.; Wiley:
New York, 1980; Vol. 9.
Mackie, J. S.; Meares, P. The diffusion of electrolytes in a cation
Literature Cited exchange resin membrane. R o c . R. SOC.London 1955, A232,
498-509.
Alexander, P. R.; Callahan, R. W. Liquid-liquid extraction and Norris, F. A. Refining and bleaching. In Bailey’s Industrial Oil and
striming of gold with microporous hollow fibers. J . Membr. Sci. Fats Products, Vol. 2,4th ed.; Swern, D., Ed.; Wiley: New York,
1987; 35: 57:71. 1982.
ADDleWhite. T. H. Fats and fattv oils. In Kirk-Othmer EncvcloDedia
- 1 Peppas, N. A.; Reinhart, C. T. Solute diffusion in swollen mem-
of Chem;cal Technology; Giayson, M., Ed.; Wiley: New kork, branes. Part I. A new theory. J . Membr. Sci. 1983,15,275-287.
1980; Vol. 9. Pons, W. A.; Eaves, P. H. Aqueous acetone extraction of cottonseed.
Beek, W. J.,; Muttzall, K. M. K. Transport Phenomena; Wiley: J. Am. Oil Chem. SOC.1967,44,460-464.
London, 1975; p 261. Prasad, R.; Sirkar, K. K. Solvent extraction with hydrophilic and
Braae, B. Degummimg and refining practices in Europe. J. Am. Oil composite membranes. AZChE J. 1987,33, 1057-1066.
Chem. SOC.1976,53, 353-357. Prasad, R.; Sirkar, K. K. Dispersion-free solvent extraction with
Callahan, R. W. Novel uses of microporous membranes: a case microporous hollow fiber modules. AZChE J . 1988,34,177-188.
study. AZChE Symp. Ser. 1988,84 (No. 261), 54-65. Sakai, K.; Takesawa, S.; Mimura, R.; Ohashi, H. Structural analysis
Dahuron, L.; Cussler, E. L. Protein extraction with hollow fibers. of hollow fiber dialysis membranes for clinical use. J. Chem. Eng.
AZChE J . 1988,34, 130-136. Jpn. 1987,20,351-356.
Dekker, M.; Koenen, P. H. M.; van’t Riet, K. Reversed micellar- Sandell, E. B. Meaning of the term separation factor. Anal. Chem.
membrane-extraction of enzymes. Znt. Chem. Eng. Symp. Ser. 1968,4,834-835.
1990, 118, 7.1-7.12. Satterfield, C. N.; Colton, C. K.; Pitcher, W. H. Restricted diffusion
DElii, N. A.; Dahuron, L.; Cussler, E. L. Liquid-liquid extractions in liquids with fine pores. AZChE J . 1973,19,62&635.
with microporous hollow fibers. J . Membr. Sci. 1986,29,309-319. Shah, K. J.; Venkatesan, T. K. Aqueous isopropyl alcohol for ex-
Faxen, H. Die Bewegung einer starren Kugel langs der Achse eines traction of free fatty acids from oils. J . Am. Oil Chem. SOC.1989,
mit zahrer Flbsigkeit gefiillten Rohres. Ark. Mat. Astron. Fys. 66, 783-787.
1923, 17, 27. Stein, W. The hydrophilization process for the separation of fatty
Keurentjes, J. T. F.; Bosklopper, Th.G.J.; van Dorp, L. J.; van’t Riet, materials. J. Am. Oil Chem. SOC.1968,45,471-474.
K. The removal of metals from edible oil by a membrane extrac- Torrey, S., Ed. Edible oils and fate, developments since 1978. Food
tion procedure. J . Am. Oil Chem. SOC. 1990,67,28-32. Technology, Rev. 57; Noyes Data Corp.: Park Ridge, 1983.
Keurentjes, J. T. F.; Linders, L. J. M.; Beverloo, W. A.; van’t Riet, Uksila, E.; Varesmaa, M.; Lehtinen, I. Separation of unsaturated
K. Membrane cascades for the separation of binary mixtures. fatty acids of soybean and linseed oils by crystallization and
Accepted for publication in Chem. Eng. Sci., 1991. subsequent liquid-liquid extraction. Acta Chim. Scand. 1966,20,
Kiani, A.; Bhave, R. R.; Sirkar, K. K. Solvent extraction with im- 1651-1657.
mobilized interfaces in a microporous hydrophobic membrane. J. Wilke, C. R.; Chang, P. Correlation of diffusion coefficients in dilute
Membr. Sci. 1984,20, 125-145. solutions. AZChE J . 1955, I, 264-270.
Kim, B. M. Membrane-based solvent extraction for selective removal Yang, M. C.; Cussler, E. L. Designing hollow fiber contactors.
and recovery of metals. J. Membr. Sci. 1984,21,5-19. AZChE J . 1986,32, 1910-1916.
Klein, E.; Holland, F. F.; Eberle, K. Comparison of experimental and Zilch, K. T. Separation of fatty acids. J. Am. Oil Chem. SOC.1979,
calculated permeability and rejection coefficients for hemodialysis 56,739A-742A.
membranes. J . Membr. Sci. 1979,5, 173-188.
Kloosterman, J.; van Wassenaar, P. D.; Bel, W. J. Membrane bior- Received for review March 1, 1991
eactors. Fat Sci. Technol. 1987,89, 592-597. Accepted October 18, 1991

Individual Longitudinal Dispersion Coefficients of Two Immiscible


Liquids in Bubble Columns
Satoru Ami* and Hidekazu Yoshizawa
Department of Chemical Engineering, University of Osaka Prefecture, Sakai, Osaka 591, Japan

The individual longitudinal dispersion coefficients of continuous and dispersed liquid phases were
measured in bubble columns with two immiscible liquids, by means of a transient-state measurement
technique. The columna were operated batchwise with respect to both liquids, over a wide range
of the relevant physical properties and average volume fraction of the dispersed liquid. The observed
individual dispersion coefficients were empirically correlated in terms of the Peclet number based
on the superficial gas velocity, as a function of the relevant system parameters.

Bubble columns may be used for operations of gas-liq- Yoshizawa, 1991; Hatzikiriakos et al., 1990a,b,).
uid-liquid systems in diverse areas (Asai and Yoshizawa, Information concerning the mixing of fluids in bubble
1991). Limited studies are available for some character- columns with immiscible liquids is important for predicting
istics in such bubble columns, including the longitudinal the concentration profiles of relevant species, or temper-
holdup distribution of the gas and dispersed liquid and ature profile, in the design or analysis of mass- and
the mean drop size and drop size distribution (Asai and heat-transfer equipment, and reactors. Hatate et al. (1975)
measured the longitudinal dispersion coefficients of dis-
* T o whom correspondence should be addressed. persed liquid for an +kerosine (dispersed liquid)-water
0888-5885/92/2631-0587$03.00/00 1992 American Chemical Society

You might also like