You are on page 1of 9

The Dehydration of Fermentative

2,3=Butanediol into Methyl Ethyl Ketone

Ai V. Tran and Robert P. Chambers


Department of Chemical Engineering, Auburn University,
Auburn, Alabama 36849
Accepted for publication February 18, 1986

A solid acid catalyst consisted of sulfonic groups covalently bound to an inorganic matrice was developed to
dehydrate 2,3-butanediol into methyl ethyl ketone. Rate
constant and apparent activation energy of the dehydration reaction were determined. The decay course of the
catalyst was a two-stage curve. The catalyst was deactivated more rapidly in the first stage than in the second
stage. The strategy of maintaining constant degree of
dehydration was employed to lengthen the lifetime of
catalyst. Treatment of the 2,3-butanediol containing fermentation broth with activated carbon greatly facilitated
the subsequent dehydration reaction.

INTRODUCTION

Recent interest in the utilization of renewable lignocelluloses has intensified research on the fermentation of xylose and glucose to 2,3-butanedi01'-~(hereafter butanediol). This is because butanediol is the
precursor of a number of compounds.6 Among those
is methyl ethyl ketone (MEK). Compared to ethanol,
MEK has a higher heat of combustion, 584.2 vs. 326.7
kcal/mol. It also gives an octane number of 96.7 when
mixed (25% volume) with gasoline.' Thus, MEK is
more effective as a liquid fuel additive than ethanol.
A two-step process from butenes is usually used to
make MEK.8 Butenes are first hydrated to give 2butanol which is then dehydrogenated over zinc or
copper based catalysts at high temperatures and low
pressures to produce MEK. The conventional conversion of fermentative butanediol into MEK is also a
two-step process. Butanediol is first recovered from
fermentation broth and purified. It is then dehydrated
over activated bentonite9 or sulfuric acidlo*' to yield
MEK. Sulfuric acid is not able to convert butanediol
in the fermentation broth directly into MEK. This is
due to the preferential reaction of sulfuric acid with
unfermented xylose in the broth to the reaction of sulfuric acid with butanediol."
In light of the above facts, the present report describes the direct dehydration of butanediol in the fermentation broth into MEK over a solid acid catalyst.
Biotechnology and Bioengineering, Vol. XXIX, Pp. 343-351 (1987)
0 1987 John Wiley & Sons, Inc.

The advantage of this method is the elimination of the


energy-intensive step of recovery and purification of
butanediol from fermentation broth prior to the subsequent dehydration reaction.

MATERIALS AND METHODS


Preparation of Solid Acid Catalysts

Alumina and silica-alumina supports were used in


the preparation of catalysts. Alumina support (35-50
mesh) was obtained from Myco, Inc. and silica-alumina supports were from Davison Chemical and
Cyanamid. The support of Davison Chemical was in
pellet form (3/16 x 3/16 in.). It was then ground, and
the 8-20 mesh fraction was retained for further use.
The support from Cyanamid was in fine form (40-60
mesh). Sulfhydryl groups were first covalently attached to the inorganic supports via silane intermediates.I2 They were then transformed into sulfonic groups
using the procedure of Backer. l 3 The transformation
reaction was carried out at 60 ? 1 and 90
1C. In
this work, the catalysts A160, D60, and C60 were those
made at 60 2 1C from alumina, Divison Chemical,
and Cyanamid supports, respectively. Similarly, A190,
D90, and C90 were the catalysts prepared at 90 2 1C
from the corresponding supports.
+_

Reactors and the Dehydration of Butanediol


Tubing Bomb

The dehydration of butanediol into MEK was first


made in a 21-cm-long, I-cm-I.D., stainless-steel tubing
bomb. After charging with A190 and butanediol (50
g/L), the tubing bomb was placed in an oil bath previously heated to a given temperature. It was then
cooled in a stream of water and filtered. The filtrate
was used for determination of butanediol and MEK.
CCC 0006-3592/87/030343-09$04.00

Batch Autoclave

A 600-ml Parr batch autoclave reactor was employed. About 13.5 g catalyst and 300 mL of 50 g/L
butanediol solution (Sigma Chemical) were used for
each run. The stirrer speed was 15 rpm. Time zero
(initial time) was the time at which the reactor reached
the given temperature. Samples were taken at intervals
after passing a sampling coil (1 m) immersed in an icewater bath. After the dehydration reaction, the reactor
was cooled and filtered. The catalyst was washed with
4 L distilled water and allowed to air-dry overnight. It
was used recurrently in three successive runs at identical conditions.

Figure 1. Scheme of the packed bed reactor system:


(1) reservoir of butanediol solution, (2) minipump, (3) pressure gauge,
(4) relief valve, (5) filters, (6) quick-fix connector, (7) oil bath, (8)
packed bed reactor, (9) coil (1 m), (10) ice-water bath, ( 1 1) fraction
collector, (12) thermometer,(13) stainless-steeltubing, (14) 200 mesh
screen, (15)glass wool, (16)glass bead (3 mm O.D.), and (17) catalyst.

Packed Bed Reactor

Figure 1 depicts the packed bed reactor system used


in this work. A minipump (Milton Roy) pumped the
aqueous butanediol solution (50 g/L) through the system. After passing 40 mL, the packed bed reactor was
immersed in an oil bath which was previously heated
to a given temperature. In order to ensure equilibrium
conditions, a forerun of 40 mL was passed before collecting the liquid products for analysis. The pressure
of the system was 300 psi (gauge pressure).

Regeneration of Catalyst

The catalyst was washed by pumping water (500 ml)


through the packed bed reactor. Water was then replaced by hydrochloric acid [5% (vh), 400 mL1. The
water-hydrochloric acid cycle was repeated three times.
Finally, the reactor was washed with water until no
trace of hydrochloric acid was detected (the indicator
was methyl orange).

Preparation of Fermentation Broth

A solution (5 L) composed of NaCl (1 g/L),


MgS047H20 (0.2 g/L), NH4CI (2 g/L), yeast extract
(6 g/L), and xylose (100 g/L) was sterilized at 120C
for 15 min in a 16-L fermentor (New Brunswick, type
SF-1 16). Inoculation (1 L) of Klebsiella pneumoniae
AU-1-d3 was added to the medium. The butanediol
fermentation was then proceeded at 32C for 15 h. The
pH was maintained at 5.4 with 5N NaOH. The source,
growth conditions, and inoculum composition of Klebsiella pneumoniae AU-l-d3 were described elsewhere.I4 The fermentation broth was withdrawn from
the fermentor, centrifuged (15 x lo3 rpm x 15 min,
at 4"C), and stored in a cold room (4C). Fermentation
products in the broth were butanediol(27.5
ethanol
(4.2 g/L), acetic acid (2.1 g/L), and acetoin (5.7 g/L).

a),

344

Treatment of Fermentation Broth

Since butanediol in the fermentation broth was not


efficiently dehydrated to MEK over the catalysts as
will be described later, two treatments were used for
the broth prior to the subsequent dehydration reaction.
In the first treatment, the broth was run through a glass
column (25 cm long x 1.2 cm I.D.) packed with Amberlite IR-120, H+ form. The treated broth was collected after passing a forerun of 35 mL. In the second
treatment, the broth was treated with activated carbon
(50 g/L, Darco grade MD 3000) at 60C for 40 min. It
was then filtered (Whatman filter paper No. 3) and
centrifuged (12 x lo3 rpm x 10 min, at 4C).
Analytical

Butanediol and MEK from the dehydration reaction


were analyzed on a Varian gas chromatograph 3700
using a 1-m glass column packed with Chromosorb 101
(60/80 mesh). The gas chromatograph was equipped
with a flame ionization detector, a Varian autosampler
5000, and a Varian integrater CDS-1 11C. Fermentation
products were analyzed on the same equipment. Sulfonic
groups of the catalyst were determined by treating
the catalyst (0.5 g) with 0.1N NaOH (50 mL) overnight
with stirring. The residual NaOH was then titrated with
0. lNHC1. pH curve of the catalyst was carried out with
a glass-electrode Horizon pH Controller 5997-20.

RESULTS AND DISCUSSION


Properties of Catalyst

Table I indicates that the catalysts made at 90 ? 1C


contained more sulfonic groups than did the corresponding ones prepared at 60 & 1C. As seen in Figure
2, the initial pH of A190, D90, and C90 was higher than

BIOTECHNOLOGY AND BIOENGINEERING, VOL. 29, FEBRUARY 1987

Table I. Sulfonic groups content and the dehydration of butanediol over various
catalysts in batch reactor.

Rate constant
[min-' g-I ( x
Catalyst
A190
Run 1
Run 2
Run 3
A160
Run 1
Run 2
Run 3
D90
Run 1
Run 2
Run 3
060
Run I
Run 2
Run 3
C90
Run 1
Run 2
Run 3
C60
Run 1
Run 2
Run 3

SO,H
(meq/g)
1.31

1.28

1.44
1.22
1.41

1.81
I .05
1.59

Overall

First
stage

Second
stage

Degree of
butanediol
dehydration (%)

32.97
18.48
13.45

12.55
9.79

26.02
16.13

100
98.6
89.2

28.47
16.55
9.80

13.73
8.08

21.04
10.84

93.4
81.5

53.37
18.67
9.13

16.84
7.03

23.11
11.10

100
97.3
91.7

44.77
18.06
7.66

14.67
5.42

22.46
9.07

100
96.5
87.0

39.85
9.62
2.61

7.92
2.24

10.65
2.74

100
94.4
86.7

32.92
11.10
4.33

10.53
3.36

11.84
5.00

that of A160, D60, and C60, respectively. Also, the pH


curves (Fig. 2) show that after ca. 0.75 mL of 0.1N
NaOH was consumed, the former catalysts had lower
pH than did the later catalysts. These results suggest
that AIW, DW, and C90 may have lower external but
higher internal sulfonic groups than do the respective
catalysts A160, D60, and C60.

100

100

90.7
84.3

A l 60
-A l 90
-. -

........ DD
_---C
_ . -c

Dehydration of Butanediol into MEK

60
90
60
90

Tubing Bomb

In order to establish the reaction conditions of butanediol dehydration, tubing bomb was used. Results
in Figure 3 indicate that the dehydration reaction was
depended on the temperature and quantity of catalyst.
As temperature increased from 150 to 220"C, ca. 100%
increase in the degree of butanediol dehydration was
obtained. On the other hand, ca. 15% increase in the
degree of butanediol dehydration was observed for a
quadruple in catalyst quantity. Thus, the effect of temperature on the butanediol dehydration was more pronounced than that of catalyst quantity. From these
data, 2 10C was arbitrarily selected as the dehydration
temperature for subsequent experiments.

12

0 . I N NaOH ( m l )
Figure 2. Titration curves with 0.1N NaOH of various catalysts.

TRAN AND CHAMBERS: DEHYDRATION OF 2,3-BUTANEDIOL

345

T e m p e r a t u r e ("C)

order in butanediol concentration. Both D90 and C90


completely dehydrated butanediol in the first run.
However, their activity was somewhat decreased in
the second and third recurrent runs as indicated by the
decreases in the reaction rate constant and degree of
butanediol dehydration of these runs (Table I). This
implies that both catalysts were deactivated. The dehydration reaction in the second and third runs followed a two-stage path (Fig. 4). Compared to the reaction rate constant of the first stage, that of the second
stage was higher (Table I). Thus, the dehydration reaction was faster in the second stage than in the first
stage. The reason of these facts will be elaborated later.
The behavior of other catalysts was similar to that
of D90 and C90. The catalysts made at 60 2 1C had
lower reaction rate constants than did the corresponding ones prepared at 90 ? 1C. This is due to the lower
sulfonic groups of the former catalysts (Table I).
The apparent activation energy of the dehydration
reaction over the catalysts was 2.9 x lo4 cal/mol, calculated from their Arrhenius plots (not shown here).
This value is lower than that found for the dehydration
of butanediol by sulfuric acid (3.6 x lo4 cal/mol)."

0
.e
0
L
V

>

JZ
W

-0
.-

-0
W

t
0
c

rn
w0

W
W
L
0,

1.o

0 "0.5

2.0

1.5

A l 9 0 (g)
Figure 3. Degree of butanediol dehydration as a function of temperature and catalyst quantity.
The empty circles show 0.5 g A190, 12 mL 50 g/L butanediol, dehydrated for 150 min in the tubing bomb; the solid circles show 12
mL 50 g/L butanediol, at 2 W C , dehydrated for 150 min in the tubing
bomb.

Packed Bed Reactor

Typical packed bed reactor profiles at 210C are illustrated in Figure 5 for the catalysts D90 and C90.
The butanediol concentration decreased to a minimal
point then increased gradually. This indicates that the
catalysts were deactivated after the maximal dehydration of butanediol occurred. Assuming that the dehydration reaction was in the steady state, i.e., the butanediol concentration after the minimal point was
unchanged, the rate constant of the reaction was evaluated from the equation: In CJCi = - k ~ ,where Ci

Batch Autoclave Reactor

Shown in Figure 4 is the typical relationship of In


C/Co with reaction time for the catalysts D90 and C90.
Parameters Co and C were butanediol concentrations
at initial and different reaction times, respectively; the
slopes of the lines were the rate constants of the reaction. Apparently, the dehydration reaction was first
0

-1

-1

-2

-2

0
\

0
c

-.

-3
1

'

40

80

120

160

200

40

80

120

160

200

T i m e (min)
Figure 4. Correlation of In (C/Co) with time for batch reactor at 210C: (A) catalyst D90, (B) catalyst C90, (1) first run,
(2) second run, and (3) third run.
346

BIOTECHNOLOGY AND BIOENGINEERING, VOL. 29, FEBRUARY 1987

c
L

50
c
0

-0

Again, the dehydration reaction rate constants of the


catalysts made at 90 2 1C were higher than those of
the counterparts made at 60 ? 1C (Table 11). Compared to the reaction rate constants measured for batch
reactor, those determined for packed bed reactor were
slightly higher except for A160. Thus, the physical
properties, i.e., the differences in the void and pore
structures of the supports used, of the catalysts have
probably affected their behaviors in packed bed reactor. The degree of butanediol dehydration of the catalysts made at 90
1C was also higher than that of
the catalysts prepared at 60 2 1"C, thus corresponding
to the higher reaction rate constants (Table 11) and
higher sulfonic groups (Table I) of the former catalysts.
Since the amount of D90 used was small (Fig. 5 ) , its
degree of butanediol dehydration was expectedly low
(Table 11).
The deactivation constant of the catalysts in packed
bed reactor was calculated using the equation k, =
ke - k d f , 1 6 where k, is the dehydration reaction rate constant at time t after the maximal dehydration point; k
is the rate constant at the maximal dehydration point;
and kd is the deactivation constant. The deactivation
constant of the catalysts made at 90 2 1C was slightly
lower than those of the counterparts made at 60 1C
(Table 11). As Figure 6 shows, the deactivation of the
catalysts was a two-stage curve. The catalysts were
deactivated faster in the first stage than in the second
stage (Table 11). These results will be discussed in detail later.

40

.-

v
0,

30
c
3

20

- 1

T i m e (min)
Figure 5. Profiles of butanediol dehydrationover different catalysts
in packed bed reactor at 210C:
(A) catalyst D90, 15.4 g, T = 7.7 rnin;
(B) catalyst C90, 27.5 g, T = 19.9 rnin;
( I ) run prior to the regeneration of catalyst;
(2) run after the first regeneration; and
(3) run after the second regeneration.

and C , are butanediol concentrations at time zero and


at the point of maximal dehydration, respectively. The
residence time, T , was computed using the feed rate
(mL/min) and void volume of the catalyst bed (mL).
The void volume was assumed to be 70% of the total
volume of catalysts bed.I5

Improvement of Catalyst Activity


Batch Autoclave Reactor

One reason of the catalyst deactivation is due to


poisons which block the active sites, i.e. sulfonic groups,
of the ~ata1yst.l~
The dehydration in batch reactor was
then proceeded with the addition of 0.5% (w/w, cata-

Table 11. Characteristics of the dehydration of butanediol over different catalysts in packed bed reactor.

Deactivation constant
[min-' ( x lo-"]

Catalyst
A190
First regeneration
Second regeneration
A160
D90
First regeneration
060
C90
First regeneration
Second regeneration
C60

Rate constant
[min-' g-l ( x
33.59
15.21
8.01
30.08
73.06
27.13
50.58
67.06
16.76
12.50
46.30

Overall

First
stage

Second
stage

Degree of butanediol
dehydration (%)

33.65
27.67
3 1.47
34.07
39.62
54.45
39.89
35.57
43.79
15.98
39.16

52.17
40.56
83.99
54.07
63.16
77.23
62.92
52.69
69.04
10.95
59.46

10.57
21.96
29.26
16.35
21.21
29.57
18.48
29.37
27.02
12.02
14.13

87.7
60.7
40.9
79.4
58.0
29.9
71.7
97.4
50.0
38.8
83.3

Conditions of the dehydration reaction are given in Figure 5.


TRAN AND CHAMBERS: DEHYDRATION OF 2.3-BUTANEDIOL

347

Time (min)
Figure 6. Deactivation curves of different catalysts used in packed bed reactor; see Figure 5 for
Codes.

lyst basis) of 0.5% platinum on alumina (40-60 mesh).


Prior to heating , the batch reactor was pressurized to
100 psi with hydrogen. As data in Tables I and I11
indicate, although the pattern of the dehydration reactions with and without platinum was the same, i.e.,
one stage in the first run and two-stage path in the
second and third recurrent runs, the rate constants of
the former reaction (with platinum) were lower than
those of the latter reaction (without platinum). These
results imply that a part of sulfonic groups was blocked
by platinum containing alumina, thus resulting in lower

rate constants of the dehydration reaction with platinum. This, in turn, suggests that the deactivation of
catalyst was not caused by poisons but rather by the
decrease in sulfonic groups. A case in point is the lower
sulfonic groups of D90 and C90 after the third run
(Table I). Due to the decrease in sulfonic groups, and
since the catalyst contained more internal than external
sulfonic groups as mentioned earlier, it may then conceive that after the first run the external sulfonic groups
were depleted at a much faster rate than were the internal sulfonic groups. This will explain the higher re-

Table 111. Rate constant and degree of the butanediol dehydration in batch reactor with the addition of 0.5% (w/w) platinum.

Rate constant
[rnin-' g - ' ( x
Overall

First
stage

Second
stage

1
2
3

35.39
15.80
7.04

13.91
6.57

17.02
7.71

2
3

34.32
13.00
5.49

12.88
3.50

13.17
7.82

1
2
3

35.22
16.45
7.82

13.71
7.39

17.75
8.06

I
2
3

23.50
10.55
4.66

9.47
3.29

11.10

Catalyst
090
Run
Run
Run
060
Run
Run
Run
C90
Run
Run
Run
C60
Run
Run
Run

348

Degree of
butanediol
dehydration (%)
100

98.2
83.5
100

96.5
84.0
100

95.9
85.7
100

5.12

BIOTECHNOLOGY AND BIOENGINEERING, VOL. 29, FEBRUARY 1987

86.5
79.7

--

50-

0,

.-0

40-

0
L

30-

0
0
C

20-

.-

-0
0)

c
0

10-

O;,

10

12

14

16

18

20

22

24

Time (hr)
Figure 7. Profiles of butandiol dehydration over different catalysts in packed bed reactor at lower reaction temperatures:
( I ) catalyst D90, 185"C, 45.9 g, 7 = 74.1 min;
(2) catalyst D60, 192"C, 44.5 g, 7 = 72.3 min;
(3) catalyst C60, I W C , 40.6 g, 7 = 83.6 min; and
(4) catalyst C90, I W C , 38.4 g, 7 = 87.7 min.

action rate constants of the second stage compared to


those of the first stage in the second and third recurrent
runs (Tables I and 111).

I +
CH3-C-C-CH3
I I

H H
(secondary
carbonium ion)

Packed Bed Reactor

H+O H

HO

--+

II

CH3-C-C-CH3

--+

H
(oxonium ion)

Regeneration of Catalyst. It is likely that the dehydration of butanediol proceeds with protons derived
from the sulfonic groups of the catalyst. Thus, the
mechanism of the reaction can be written as follows:

HO

OH

I I
CH3-C-C-CH3
1 1

HO

H'

-+

H H
(butanediol)

OH+*

I I
CH3-C-C-~~3
I 1
H

---

The protons, in ideal state, will be recycled in the


dehydration reaction. Since the catalyst was deactivated as proven above, the sulfonic groups might lose

Table IV. Rate constant, deactivation constant, and degree of butanediol dehydration of various catalysts
in packed bed reactor.

Deactivation constant
[rnin-I ( x 10-91
Catalyst

D90
D60
C90
C60

Rate constant
[min-' g-I ( X
2.35
3.18
8.14
6.78

Overall

First
stage

Second
stage

Degree of butanediol
dehydration (%)

2.12
3.85
4.60
4.67

6.27
11.97
9.84
7.13

I .67
2.89
3.45
3.82

55.0
63.2
93.6
89.9

Conditions of the dehydration reaction are given in Figure 7.


TRAN AND CHAMBERS: DEHYDRATION OF 2,BBUTANEDIOL

349

of the dehydration reaction


Table V. Rate constant [min-l g-' ( X
of butanediol in simulated fermentation broth over the catalyst D90 in batch
reactor using conditions of Figure 4.
Feed composition

Run I

Run 2

Run 3

Butanediol (50 g/L)


Simulated fermentation broth (SFB)"
SFB without yeast extract

53.37
12.58
53.31

18.67
3.50
17.31

9.13
0.03
9.31

a Composition of SBF (g/L): ethanol, 11.5; acetic acid, 4.5; acetoin, 2.3;
butanediol, 30; NaCI, 1; MgSO4.7H2O,0.2; NH,CI, 2; yeast extract, 6; and
xylose, 100.

their protons. HC1 was then used to regenerate the


catalyst. Figure 5 illustrates the profiles of the dehydration over the regenerated catalysts. The regeneration process did not improve the catalyst activity. The
rate constant and degree of butanediol dehydration of
the reaction over the regenerated catalyst were lower
than those of the reaction over the fresh catalyst (Table
11). These results imply that the deactivation of catalyst
was due to the loss of sulfonic groups in the reaction,
thus supporting the above conclusion on the catalyst
deactivation in batch autoclave reactor.
The above suggestion that during the dehydration
reaction the catalyst lost more external than internal
sulfonic groups can also be used to account for the
two-stage deactivation curve of the catalyst (Fig. 6).
It may envisage that the dehydration reaction in packed
bed reactor proceeded first with the external, then with
the internal sulfonic groups. Consequently, the deactivation of the catalyst was a two-stage curve. Since
the remaining external sulfonic groups were lower as
already suggested, it is understandable that the deactivation constant in the second stage was lower than
that of the first stage.
Maintenance of Constant Degree of Butanediol Dehydration. One of the strategies to lengthen the catalyst
lifetime in packed bed reactor is to maintain a constant
degree of conver~ion.'~
This can be done by either
varying the reaction temperature with time while keeping the feed rate constant or changing the feed rate
Table VI.

Weight
(g)

Temperature
("C)

D90
D90
D90

35.2
35.5
35.3

185
185
185

C90
C90
C90

38.4
35.4
34.7

190
190

C90

35.2

190

350

Dehydration of Fermentative Butanediol

The catalyst D90 was first employed to examine the


effect of different components in the simulated fermentation broth on the dehydration of butanediol in
batch reactor. Data in Table V indicate that xylose,
mineral salts, and fermentation products, i.e. butanediol, ethanol, acetic acid, and acetoin, did not hinder
the dehydration reaction. Yeast extract, however, inhibited the dehydration of butanediol, presumably
through the blockage of sulfonic groups. This is furthered by the fact that butanediol in the actual fer-

Dehydration of butanediol in actual fermentation broth in packed bed reactor using conditions of Figure 7.

Catalyst

while holding the reaction temperature constant. The


latter method was selected for the present catalyst. The
reaction temperature, however, was maintained at ca.
190C. This is because the temperature of the oil bath
could not be raised to 210C. This fact, in turn, demanded an increase in the catalyst quantity. The profiles of this experiment is shown in Figure 7. The degrees of butanediol dehydration were almost the same
for the reactions at 210 and 190C (Tables I1 and IV).
However, the lifetime of the catalyst was maintained
longer, 24 h vs. 6 h (Figs. 5 and 7), although this resulted in lower deactivation constant and reaction rate
constant. Also, the deactivation course of the catalyst
was unchanged. That is the catalyst was deactivated
more rapidly in the first stage than in the second stage
(Table IV).

190

Feed

Rate constant
[min-' g-' ( x lo-")]

Degree of butanediol
dehydration (%)

2.84
I .27
2.71

57.5
21.2
44.4

8.14
3.57
3.96

93.6
60.6
65.7

7.01

80.7

butanediol (27.5 g/L)


actual fermentation broth"
actual fermentation broth
treated with activated carbon
butanediol (50 g/L)
actual fermentation brotha
actual fermentation broth
treated with IR-120, H'
actual fermentation broth
treated with activated carbon

See the Materials and Methods section.


BIOTECHNOLOGY AND BIOENGINEERING, VOL. 29, FEBRUARY 1987

mentation broth could not be effectively dehydrated


(Table VI). Again, the color bodies deposited on the
catalyst were the cause. Removal of the color bodies
by treating the actual fermentation broth with activated
carbon greatly facilitated the subsequent dehydration
reaction. On the other hand, treatment with Amberlite
IR-120, H did not affect much the dehydration over
the catalyst of the treated broth (Table VI). Similar
results were also obtained for the treatments with Amberlite IR-45,OH-, and trichloroacetic acid.
+

This research was financed by grants from the Advanced


Manufacturing Technology Center and the Pulp and Paper
Research and Engineering Center, Auburn University.

References
N. B. Jansen and G. T. Tsao, in Advances in Biochemical EnRinreringlBiotechnology, A. Fiechter, Ed. (Springer-Verlag, New
York, 19831, pp. 85-99.
N. B. Jansen, M. C. Flickinger, and G. T. Tsao, Biotechnol.
Bioeng., 26, 362 (1984).
V. M. Laube, D. Groleau, and S. M. Martin, Biotechnol. Lett.,
6 , 535 (1984).
J. M. Sablayrolles and G. Goma, Biotechnol. Bioeng., 26, 148
(1984).

5. A. Willetts, Biotechnol. Lett., 6 , 263 (1984).


6. G. A. Ledingham and A. C. Neish in Industrial Fermentations,
L. A. Underkofler and R. J. Hickey, Eds. (Chemical Publishing
Co., New York, 1954), Vol. I , pp. 27-93.
7. F. Baronnet, M.Niclausa, A. Ahmed, R. Vischnievski, and L.
Charpenet, French Patent, French Demande 2367110 (May 5,
1978).
8. Kirk-Othmer, Encyclopedia of Chemical Technology, 3rd ed.
(Wiley, New York, 1981), Vol. 13, p. 905.
9. A. N. Bourns and R. V. V. Nicholls, Can. J. Research, 25B,
81 (1947).
10. A. C. Neish, V. C. Haskell, and F. J. MacDonald, Can. J.
Research, 23B, 281 (1945).
11. R. R. Emerson, MS thesis, Purdue University, L.afayette, IN,
1981.
12. R. P. Chambers, G. A. Swan, E . M. Walle, W. Cohen, and W.
H. Baricos, in Immobilized Enzyme Technology, H. H Weetall
and S. Suzuki, Eds. (Plenum, New York, 1979, pp. 199-223.
13. H. J. Backer, Rec. Trav. Chim., 54, 215 (1935).
14. A. V. Tran and R. P. Chambers, Appl. Microbiol. Biotechnol.,
23, 191 (1986).
15. J. M. Smith, Chemical Engineering Kinetics, 3rd ed. (McGrawHill, New York, 1981), pp. 327-356.
16. 0. Levenspiel, Chemical Reaction Engineering, 2nd ed. (Wiley,
New York, 1972), p. 544.
17. J. R. Gonzalez-Velasco, M. A. Gutierrez-Ortiz, J. I. GutierrezOrtis, and A. Romeo, Chem. Eng. J., 28, 13 (1984).

TRAN AND CHAMBERS: DEHYDRATION OF 2.3-BUTANEDIOL

351

You might also like