Sei sulla pagina 1di 14

Energy in Agriculture, 6 (1987) 63-76 63

Elsevier Science Publishers B.V., Amsterdam - - Printed in The Netherlands

Dehydration of Aqueous Ethanol

B. TANAKA and L. OTTEN


School of Engineering, Ontario Agricultural College, University of Guelph, Guelph, Ont. NIG
2W1 (Canada)
(Accepted 11 September 1986)

ABSTRACT

Tanaka, B. and Otten, L., 1987. Dehydration of aqueous ethanol. Energy Agric., 6: 63-76.

Upgrading of aqueous alcohol to anhydrous ethanol may be accomplished using a packed bed
adsorption process, in which cracked grain corn is the adsorbent. A 0.35 m diameter by 3.0 m
dehydration column was designed and constructed on the basis of data obtained in a series 'of
bench-scale experiments.
The results demonstrated that grain corn would upgrade 91% ethanol to 99%-plus at a rate of
about 0.20 L/min. The capacity of the corn bed in the prototype ranged from 7.6 to 10.5 mL/kg
of bed, which was lower than expected from the bench-scale experiments. The difference was
attributed to the significant thermal effect of the heat of adsorption, which caused higher bed
temperatures.
The performance of the prototype was modelled mathematically using a one-dimensional dis-
persive-convective description of the bed. The model was observed to fit the experimental data
well in the regions where heat effects were not pronounced (where C/Co<0.5), and showed a
systematic departure in the non-isothermal regions.
The estimated energy consumption for the dehydration and regeneration cycles was observed
to be lower than that normally associated with azeotropic distillation. Aside from the energy
advantage, an adsorption system is easier and safer to operate than azeotropic distillation.

INTRODUCTION

Concern for the reliability of petroleum supplies has resulted in an interest


in the use of ethanol as a gasoline extender. Ethanol is of interest primarily
because it can be derived from a renewable resource by the fermentation pro-
cess. The production of ethanol consists of a fermentation step that produces
a low grade ethanol-water mixture which contains 6-12% ethanol by volume.
Ethanol is separated from this mixture by a two-step distillation which yields
anhydrous ethanol that can be used in gasohol. The conventional two-step
distillation proceeds with the distillation of the dilute ethanol-water mixture
to an azeotropic concentration of 95.6% ethanol by weight. This stage is fol-
lowed by the distillation of the azeotrope in the presence of a third component
which permits the recovery of 100% ethanol. The conventional separation

0167-5826/87/$03.50 © 1987 Elsevier Science Publishers B.V.


64

methods are very effective but the two-stage distillation is energy intensive.
Most of the energy is used to distill above 85% ethanol. This is due to the shape
of the vapour-liquid equilibrium curve, which indicates that at near azeotropic
concentrations any increase in ethanol concentration is accompanied by a dis-
proportionate increase in energy consumption. Eakin et al. (1981) reported
that the second step required 34% of the total energy input.
For ethanol to become a viable energy source the energy input must be
reduced. As most of the energy is required above an ethanol concentration of
85%, it would appear that this is the area where the greatest energy savings
could be accomplished. Eakin et al. (1981) performed a preliminary study to
evaluate various alternate methods of separating ethanol and water. One of the
recommendations of their study was for further laboratory investigations to be
carried out on the dehydration of aqueous ethanol using fermentable grains as
an adsorbent material which was originally suggested by Ladisch and Dyck
(1979).
This paper describes an experimental pilot-scale process that utilized the
concept of adsorption of water vapour by fermentable maize to dehydrate
aqueous ethanol. The performance characteristics of the process equipment
were evaluated and were subsequently inserted in a model for packed bed
adsorber columns.

LITERATURE REVIEW

Dehydration of azeotropic ethanol vapour by adsorption of water onto dry


CaO or K2CO3 has been reported in the literature since the 1930's (Ellis, 1937).
Although this procedure has been a standard laboratory technique for dehy-
dration for more than 50 years, the use of organic materials as adsorbents
appears to be a recent development.
The adsorption properties of water on starch in an ethanol-water-starch
system are known at temperatures of 20°C and vicinity (Chung and Pfost,
1967; Duprat and Guilbot, 1975). Water vapour isotherms were constructed
by Duprat and Guilbot for the ethanol-water-potato starch system. Using Bru-
nauer's classification of isotherms ( Adamson, 1967, p. 20), the isotherms were
observed to demonstrate the typical type II shape. These isotherms have three
distinct zones in which the adsorption process goes from very favourable at
low water activities to unfavourable at high water activities. Duprat and Guil-
bot observed that in the initial favourable region where the starch was very
dry, the adsorption isotherms for the ethanol-water mixture are indistinguish-
able from those of pure water. As the adsorbent becomes more saturated with
adsorbed water molecules, ethanol was observed to be taken up by the adsor-
bent as well. Duprat and Guilbot suggested that this effect was due to the
ability of sorbed water to dissolve ethanol in both the liquid and vapour phase.
Ladisch and Dyck (1979) proposed the use of cereal grain as an adsorbent
65

for dehydrating fuel alcohol. They tried a variety of substances as adsorbents,


including cornstarch, sugar, cellulose, grain corn and corn residue, as well as
known inorganic adsorbents such as CaO, NaOH and CaSO. The results of
these trials indicated that cellulose, cornstarch and grain corn gave similar
performance to CaO; that is, the organic adsorbents were capable of dehydrat-
ing ethanol to a purity of greater than 198-proof (99% alcohol by volume). In
addition, regeneration of these organic adsorbents was observed to require far
less energy than CaO. Ladisch and Dyck report that regeneration of the cel-
lulose adsorbent required 430 kJ/kg anhydrous ethanol produced, while CaO
adsorbent required 900 kJ/kg anhydrous ethanol produced. This difference
arises because the temperature for regenerating CaO (160-170°C) is higher
than that for regenerating cellulosic materials (80-100 ° C).
The research group at Purdue University conducted further investigations
into the adsorption characteristics of adsorbent materials. Hong et al. (1982)
developed a modified gas chromatographic elution method in which the sta-
tionary phase in the gas chromatograph was replaced by the adsorbent mate-
rial. Elution profiles were presented for corn starch, potato starch, Avicel
(microcrystalline cellulose, FMC Corp., Philadelphia, PA), xylan, cornmeal,
corn residue, wheat straw, bagasse, silica gel and protein. The profiles were
marked by two major peaks, one for ethanol, the other for water. The relative
separation of these peaks determined the suitability of the adsorbent to sepa-
rate ethanol from water. Hong et al. observed that the adsorptive capacity of
cornmeal is primarily due to the complex polysaccharide components, namely
starch and xylan. The protein component was noted to have a lower efficiency
of separation.
The Purdue team has reported also that bench-scale experiments had dem-
onstrated that corn grits would selectively remove water from vapour mixtures
containing from 50 to 95% ethanol (Bienkowski et al., 1983 ). The focus of this
research was the problems associated with regenerating a larger deep bed (41
mm diameter X 580 mm ). They observed that the bed was readily regenerated
with heated nitrogen gas and that the regeneration process had no effect upon
the adsorbent capacity. In addition, a few adsorption isotherms were developed
for ground corn for temperatures above 70 ° C.
The most recent publication by the Purdue team described the problems of
scaling up adsorption columns ( Ladish et al., 1984). Larger bore columns were
observed to be more adiabatic in nature than the small bore bench-scale col-
umns. This is due to the lower effective heat transfer area to column volume
ratio of larger columns. They report testing a multi-bed arrangement where
the beds were regenerated by heated air. The capacity of these adsorbers, which
were packed with 18 kg of corn grits, were reported to be from 3 to 12 kg of
product per h. The overall energy requirements were 1760 kJ/kg of anhydrous
product, having begun with a liquid feed of 92% ethanol by volume at ambient
temperatures.
66

COOLING WATER
VAPOUR SAMPLING

CONDENSER

JACKET

2oo= COLUMN

OIL HEATER

ALCOHOL FEED

FLASK TANK I I~
.t__lll SOALES

HEATERS BOILER

Fig. 1. Schematic diagram of the prototype adsorption apparatus.

Robertson et al. (1983) have developed a similar separation system for the
U.S. Department of Agriculture. Their approach differed from the Purdue group
in that they accomplished the dehydration at room temperatures. Cornmeal
was not as water vapour elective at these temperatures and significant amounts
of ethanol were adsorbed by the corn. To eliminate these losses the authors
suggested that the adsorbent be used as feedstock for subsequent fermentation
batches.
Work closely paralleling that at Purdue was initiated by Agriculture Canada
in 1982 (Otten, 1985 ). This work had as its overall objective the development
of a prototype dehydration system for a farm-scale fuel ethanol plant. Prelim-
inary bench-scale experiments were performed to determine the dehydration
properties of bentonite, silicagel, woodshavings, rice and grain corn.
The main conclusion of the laboratory phase was that cracked grain corn
exhibited the best adsorption characteristics. It was therefore decided to design
and build a prototype adsorber capable of removing 2.5 L of water per cycle.
The results of these experiments are presented in this paper.

MATERIALS AND METHODS

A schematic diagram of the prototype adsorption equipment appears in Fig. 1.


The adsorption column was designed using data from the bench-scale experi-
ments. As a rule of thumb, adsorption columns are designed to provide a bed
that is 3-5 mass transfer zones deep ( Kovach, 1979). The mass transfer zone,
MTZ, of the bench-scale column was calculated using the formula:
67

M T Z = V At (1)
where V is the velocity of the concentration wave in the adsorber, and At is the
time period between the appearance of the first detectable outlet water con-
centration and the point when this concentration approaches the inlet concen-
tration. For an operating temperature of 85 ° C, the M T Z was calculated to be
approximately 0.85 m, so that a bed of 3 to 5 MTZ's should be 2.55 m to 4.25 m
in depth. For reasons of economy and to facilitate fabrication a height of 2.74 m
was selected.
The diameter of the column was determined by the adsorbent capacity and
density. The results of the laboratory phase had indicated that the operating
conditions were to be as follows: (a) cracked corn bulk density of 625 kg/m 3
and particle density of 1280 kg/m3; (b) moisture content of less than 4% d.b.
and, (c) a column temperature of approximately 85-95°C; and (d) a bed
porosity of 0.488. For these conditions the adsorbent capacity was observed to
be between 21 and 15 mL of water adsorbed per kg of dry adsorbent. Given a
desired dehydration capacity of 2.5 L of water per adsorbent bed, a mass of
120-170 kg of cracked corn is required. Since the bed depth was set at 3 MTZ's,
the diameter must be in the range of 0.28-0.34 m. Therefore, a standard 0.33 m
diameter pipe was used as the adsorption column.
The column temperature was maintained with an external heating jacket
through which light hydraulic oil circulated at a temperature of 87°C and a
flowrate of 2050 L/h. The jacket contained eight vertical baffles which pro-
moted complete circulation of the heating fluid around the column circumfer-
ence. The outer wall of the jacket was insulated with 50 mm of fibreglas
insulation.
The vapour feed to the column was provided via a boiling vessel which was
attached directly to the base of the adsorption column. Electric heaters with
an output of 6.0 kW were used to supply the energy to vapourize the aqueous
ethanol. The heaters were connected to a variac which was used to control the
rate of vapourization at 12-15 L/h.
The condenser was designed as an eight pass tube-and-shell heat exchanger
in which ethanol vapour enter the shell side through a 50 mm insulated pipe.
Cooling water was passed through finned copper tubing such that the alcohol
product could be subceoled.
The adsorber was regenerated using heated air. An airstream of 66 L / s was
provided by a 1.1-kW centrifugal fan and heated with a 6-kW dryer element.
Heated air passed in a counter current direction to the adsorption wave; i.e.,
desorption proceeded from the top down.

PROCEDURE

Prior to operation of the adsorber, it was necessary to ensure that the adsor-
bent was below 4% d.b. moisture content. Therefore the column was desorbed
68

using heated air at flow rates between 66 and 94 L / s and temperatures between
91 and 109 ° C. A thermocouple was placed at the inlet to the column to record
the drying air temperature, while at the outlet two thermocouples were arranged
to record the wet and dry bulb temperatures of the exhaust air. This permitted
calculation of the absolute humidity of the exhaust, which in turn gave an
indication of the dryness of the bed.
Once the column temperature was established, aqueous ethanol was pumped
into the boiling vessel and vapourised. The vapours passed through the adsor-
bent bed and were dehydrated by the cracked corn adsorbent. The water con-
tent of the effluent was measured by the Karl Fischer titration method. A small
portion of the vapour effluent was diverted from the top of the column to a
glass laboratory condenser with a sealed 200-mL flask, from which samples
were withdrawn with a syringe.
The temperature profile of the adsorber bed was monitored with a series of
copper-constantan thermocouples placed at depths of 100 mm, 200 mm, 300
mm, 0.5 m, 1.0 m and 2.0 m below the bed surface. In addition, the temperature
of the column wall and of the hydraulic heating fluid were monitored, all at 5-
min intervals.

EXPERIMENTAL DESIGN

The experimental procedure was divided into two parts. The first part sought
to determine the effects of scaling up of the adsorption apparatus from a bench-
scale to a farm-scale prototype. The second part was concerned with the acqui-
sition of operating parameters which were then inserted into a mathematical
model of the adsorption process.
The experiments of scale were performed with a bed depth of 2.74 m. The
standard adsorption procedure as outlined above was followed until the output
product was noted to contain more than 1% water by volume. At this point the
dehydration process was terminated and the adsorbent bed regenerated. Five
replications were performed in order to determine an average breakthrough
capacity for the adsorbent. Further experiments were performed in which
field conditions were simulated by operating the column in a cyclic manner.
Two sets of experiments were performed in which the adsorber bed was run to
the onset of breakthrough, immediately regenerated and run again to the
breakthrough point.
The model is a material balance coupled with the appropriate initial and
boundary conditions. The analysis is based on the continuous flow of a carrier
fluid through a packed column and a known equilibrium isotherm to describe
the sorption kinetics. If it is assumed that only one-dimensional dispersive and
convective mechanisms in the flowing phase are important, the equation
becomes (Yang and Tsao, 1982 ) :
69

OC 02C OC
R--~ = Dz Oz2 V 0--~ (2)

where C is the water vapour concentration (% by volume); Dz longitudinal


dispersion coefficient ( m2/s ) ; t time ( s ) ; V velocity of the adsorbate ( m/s ) ; z
distance from inlet ( m ) ; R = l + p K / e = r e t a r d a t i o n factor assuming a first-
order reaction; K adsorption rate constant (m3/kg) ; e void fraction of the bed;
and p bulk density of the adsorbent (kg/m 3).
The differential model is subject to the following boundary and initial
conditions:
(1) Flux type boundary condition at the bed inlet (Dankwertz condition) :

(3)
- Dz OC +vc =
( VCo O<t<tb
~:o 0 t > tb

(2) Constant flux at the end of the bed (where z = L) :


OC
--(L,t) =0 (4)
Oz
(3) Initial condition:
C(z, 0) = 0 (5)
There is some question about the appropriateness of the second boundary
condition, as this condition can only apply if true plug flow occurs. This is not
the case for the absorber column but there is no simple condition to describe
the flux at that boundary. Furthermore, D. Elrick (personal communciation,
1984) has used it successfully in analyzing one-dimensional convective-dis-
persive models in soil-water systems,
An analytical solution of the above set of equations has been presented by
Van Genuchten and Alves (1982). The solution is a relatively complex expres-
sion involving inverse error functions and exponential terms and relates a
dimensionless concentration ratio, C/Co, to time and bed distance. The axial
dispersion coefficient D, the retardation factor R, and the adsorbate velocity
V, are the process parameters contained in the various terms of the solution.
By evaluating the concentration ratio at the bed outlet, the solution becomes
a function of time only and represents the breakthrough data of the experi-
ments. Furthermore, the retardation factor may be calculated from the bed
conditions and K, which is the slope of the adsorption isotherms for water on
corn particles at 4% (d.b.) moisture content. The isotherm data reported by
Ladisch et al. (1984) was used to determine that R = 1.22 for the bed of cracked
corn.
70

TABLE 1

Summaryof dehydration capacity experiments using a 2.74 m deep, 146.6 kg bed and an effluent
concentration of 1% water as the end point

Exp. Boiling Influent w a t e r Adsorbed Breakthrough Product


rate concentration (mL/kg time (L/min)
(L/min) (% by volume) of bed) (min)
1 0.22 7.84 9.7 92 0.21
2 0.22 7.77 9.7 95 0.20
3 0.24 7.50 8.5 80 0.23
4 0.21 7.79 3.7 45 0.20
5 0.25 7.89 10.5 90 0.23

Therefore, to complete the model solution, breakthrough curves for beds of


0.95, 1.83 and 2.74 were obtained. The process parameters D and V were cla-
culated from the curves using the method of moments as described by Leven-
spiel (1972) and Turner (1972).

RESULTS AND DISCUSSION

Scale up effects

Dehydration capacity. Five replications were performed under similar condi-


tions to determine the effects of scale up. The most important property inves-
tigated was the capacity of adsorbent at the point of breakthrough. Table 1
contains the results and conditions of the five replications. Ignoring the anom-
alous fourth entry, the average adsorbent capacity was 9.6 m L / k g of dry corn
with a standard deviation of 0.7 mL/kg. There was no observable decrease in
the adsorbent capacity over the period in which the five trials were performed.
However, the capacity was lower t h a n expected and was between 54% and 36%
lower than the capacity observed in the bench-scale apparatus (Otten, 1985),
even though the vapor velocities were similar.
The reduced capacities are attributed to the different temperature profile of
the prototype bed. In the bench-scale apparatus the operating temperatures
were observed to be on average 10-15 °C less t h a n those observed in the pro-
totype. The temperature difference is due in part to the scaling up of the equip-
ment, as the larger bore column proved to be much less isothermal t h a n the
smaller bore bench-scale column. Since the smaller bore column has a higher
effective surface area to volume ratio, the heat removed by the jacket fluid was
greater. This result corroborates those reported by Ladisch et al. (1984);that
is, the larger scale columns behaved in a more adiabatic manner than the bench-
scale equipment.
The apparent decrease in adsorbent capacity with an increase in bed tem-
71

120

TEMPERATURE HISTORY
115
- run

110

A
C~
105
w

u.I
t~r

I-- 100
n-

THERMOCOUPLE DEPTH
w
~- 95 100 mm

90

85 I I I I I I
0 20 40 60 80 100 120 140

TIME IN MINUTES

Fig. 2. Bed temperature history of experiment 11 at 100 m m from the inlet.

perature is consistent with adsorption theory. The region of favourable uptake


in a type II adsorption isotherm decreases as temperature increases, making
the adsorption reaction less probable. Ladisch et al. (1984) noted similar results.
Using the criteria developed by Pan and Basmadjian (1970) they determined
that under certain operating conditions a pure thermal wave will precede the
adsorption front, heating up the adsorbent material. If the velocity of the car-
rier fluid is greater than the velocity of the adsorption front, the heat genera-
tion by adsorption will be swept ahead of the adsorption reaction. This causes
the adsorbent temperature to increase, which in turn lowers the effective water
equilibrium capacity of the adsorbent. In addition, Garg and Ausikaitis (1983)
noted that the MTZ elongated significantly under these conditions, which
caused premature breakthrough and therefore decreased capacity.
One positive aspect was noted. Since the thermal wave was observed to pre-
cede the concentration front, a suitably located temperature sensor could be
used to determine when the bed should be regenerated. Figure 2 illustrates the
change in temperature in the bed.

Cyclic operation of adsorber. Two sets of two experiments were performed in


which the column was operated as if it were in place at an ethanol production
facility. The column was regenerated for 90 min immediately after an outlet
concentration of 1% water (by volume) was achieved, after which it was run
again. All four experiments were completed sequentially.
72

TABLE 2

Summary of cyclic adsorption-desorption experiments with a 2.74 m deep bed, initial column
temperature 72 ° C

Exp. # Boiling Initial water Water Breakthrough Product


rate concentration, Co adsorbed time (L/min)
(L/min) ( % by volume) (mL/kg) (min)

6A 0.24 8.97 8.7 68 0.22


6B 0.23 8.97 8.9 55 0.27
7A 0.22 7.94 7.6 58 0.24
7B 0.22 7.94 7.6 58 0.24

The summary of the results is found in Table 2. The boiling or evaporation


rate and the initial feed concentration were kept in the same range as in the
dehydration experiments of Table 1. Although the adsorption capacity of the
bed averaged about 8.2 mL/kg of dry corn as compared to 9.6 mL/kg for the
first group of rexperiments, the endpoints were reached earlier so that the
ethanol production rate was 0.24 L/min as compared to 0.22 L/min. Therefore,
the cyclic operation of the column did not reduce the rate of alcohol production.
The regeneration results illustrated in Fig. 3 indicate that the desorption
rate is initially high but reduced to nearly zero after 60-80 min. The relatively
rapid regeneration process demonstrates that the water on the adsorbent is

.08
REG ENERATION CYCLE

:~ .06

.04

.02

1'-
0 I I I 1
0 20 40 60 80 100
TIME IN MINUTES

Fig. 3. Regeneration curve produced during desorption of bed with air at 100 °C and 94 L/s.
73

easily removed by convective mass transport. This suggests that the water mol-
ecules are physically adsorbed by relatively weak attractive forces on and in
the corn particles.
The cyclic operation of the column also presented an opportunity to deter-
mine the energy requirements of the process. The electric heaters employed in
the boiling vessel and the dryer element used to heat the regeneration air-
stream were not selected with efficiency in mind. Nevertheless the energy con-
sumed proved to be low in comparison to the conventional dual distillation
system. The overall energy required in producing anhydrous ethanol was 2880
kJ/kg anhydrous product, starting from an initial ethanol feed concentration
of 92 % at ambient temperatures. Ladisch et al. (1984) reported a figure of 1760
kJ/kg anhydrous product in a multibed arrangement with similar feed concen-
trations. Even at 2880 kJ/kg anhydrous product, there is a significant savings
over the azeotropic distillation step, which was reported to consume 3320 kJ/kg
(Eakin et al., 1981) at an initial feed of 95% ethanol.

Comparison of model and process performance

Adsorption experiments performed on bed depths of 0.95, 1.83 and 2.74 m


were continued until complete bed exhaustion was achieved. The concentra-
tion history was recorded as C/Co and plotted against time. An example of a
representative breakthrough curve appears as Fig. 4.

BREAKTHROUGH CURVE
z .8 run //11
o
<C
n-
I-
Z
LLI
Z .6
o o o o

~
z .4
0

Q
.2

I I
0 20 40 60 80 100 120 140

TIME IN MINUTES

Fig. 4. Representative breakthrough curve C/Co versus time ( C o = 9 . 1 4 % water, L=1.83 m,


T=90-100°C).
74
1.0

DATA
.8 _ MODEL o o
~Z ° 0
o 0 0

~ o
o

O
Z
O
O O = 0.0077 m2/min
u)
co .4 V = 0.0349 m/rain

z R = 1.22
_o

O o I 1 I
0 20 40 60 80 100

TIME IN MINUTES

Fig. 5. Model b r e a k t h r o u g h curve a n d experimental p o i n t s (D = 0.0077 rn2/min, V = 0.0349 m / m i n ,


R-- 1.22).

These data were used to determine the dispersion coefficient and the velocity
of the concentration wavefront by the method of moments. The results from
this analysis were inserted into the analytical solution of the model equation.
Figure 5 is a typical example of the results generated from the model. The
model was observed to give a good fit for the area where C/Cowas less than 0.5.
The deviation above this region is due to the increasingly adiabatic nature of
the bed with time. The analysis of Garg and Ausikaitis (1983) indicated that
the M T Z would elongate as the adsorbent temperature increased. This is borne
out in our results by the decreasing slope, hence longer MTZ, of the break-
through curve.
The model was based on the assumption that the process is isothermal; how-
ever, it was apparent that a thermal wave passed through the bed prior to the
adsorption wave. The effects were observed to be cumulative in that the bed
temperatures increased with axial position. Therefore, the longer the process
proceeded, the more adiabatic in nature the column became. Since a systematic
departure from isothermal conditions was observed, deviations should become
more pronounced with time and depth. The results shown in Fig. 5 show this
systematic deviation to be the case.
The results in Table 3 indicate that the dispersion coefficient increased with
bed depth. It is interesting to note that the velocity of the wavefronts are much
lower than the vapor velocity of about 1.5 m/min.
75

CONCLUSIONS

The prototype adsorber developed in this study demonstrated that cracked


corn is an attractive medium by which ethanol may be dehydrated. The exper-
iments of scale demonstrated that the adsorbing capacity of cracked corn was
unchanged after repeated use; however, the capacity was noted to be from 54
to 38% less than that observed in the bench scale apparatus. This loss in capac-
ity is attributed to the heat of adsorption which is retained within the adsorber
column, thereby raising the bed temperature and lowering the adsorbent
capacity.
Regeneration results indicated that the adsorbed water is easily removed by
heated air. The desorption process is comparable in length to that of the time
to reach an effluent concentration of 1% water. This would permit continuous
process operation with two or three columns to dehydrate about 13-15 L of
92% ethanol/h. The parallel column arrangement would also allow use of the
heat extracted from the active bed to assist in regenerating the other columns.
Thus increasing the energy savings.
It should also be noted that the same bed was used for all the experiments
so that it is not necessary to replace the bed frequently. In fact, in some bench-
scale experiments the same bed was used for 30 runs without adverse effects
on the performance. Of course, if one were to replace the bed the corn could be
used in the next fermentation batch.
The energy consumption of the process was noted to be less than that asso-
ciated with conventional azeotropic distillation. This is remarkable in view of
the fact that great inefficiencies existed in the desorption heater and the boil-
ing vessel heaters.
The model used to predict the column performance was observed to fit the
experimental breakthrough curves in the regions where C/Cowas less than 0.5.
Significant departures from the model were observed above this point which
were due to the increasingly adiabatic nature of the column's operation. Since
we are normally only interested in effluent concentrations of 1% water or less,

TABLE 3

Experimental model parameters obtained from complete adsorption experiments

Run Boiling Influent Initial Bed Dispersion Concentration Retardation


rate water bed height, coefficient, wave velocity factor,
(L/min) concentration temperature L D V R
(% by volume) (°C) (m) (m~/min) (m/min)

8 0.43 7.94 80 0.95 8.7 × 10 :~ 4.81 × 10 " 1.22


9 0.33 5.80 86 0.95 4.6X 10 :~ 4.20X 10 ~' 1.22
10 0.26 5.23 85 1.83 9.7 X 10 :~ 3.31 × 10 -' 1.22
11 0.24 9.14 87 1.83 7.7 X 10 :~ 3.49X 10 " 1.22
12 0.29 7.23 84 2.74 3.5 X 10 ~ 2.49 X 10 " 1.22
76

t h e m o d e l m a y be u s e d to s i m u l a t e t h e p e r f o r m a n c e of beds under these


conditions.

REFERENCES

Adamson, A.W., 1967. Physical Chemistry of Surfaces. Wiley-Interscience, New York.


Bienkowski, P.R., Voloch, M., Ladisch, M.R. and Tsao, G.T., 1983. Non-isothermal adsorption
of water vapour on corn. ASAE Pap. 83-3567, American Society of Agricultural Engineers, St.
Joseph, MI, 27 pp.
Chung, D.S. and Pfost, H., 1967. Adsorption and desorption of water vapour by cereal grains and
their products. Part I: Heat and free energy changes of adsorption and desorption. Trans.
ASAE, 10: 549-557.
Duprat, M.F. and Guilbot, A., 1975. Solvent versus non-solvent water in starch-alcohol-water
system. In: R.B. Duckworth (Editor), Water Relations of Foods. Academic Press, London.
Eakin, D.E., Donvan, J.M., Cysewslie, G.R., Petty, S.E. and Maxham, J.V., 1981. Preliminary
evaluation of alternative ethanol/water separation processes. Report prepared for U.S. Depart-
ment of Energy, Contract DE-AC06-76RLD 1830. Pacific Northwest Laboratory, Battelle
Memorial Institute, Richland, WA.
Ellis, C., 1937. The Chemistry of Petroleum Derivatives, Vol. 2. Reinhold, New York.
Garg, D.R. and Ausikaitis, J.P., 1983. Molecular sieve dehydration cycle for high water content
streams. Chem. Eng. Prog., 79 ( 4 ) : 60-64.
Hong, J., Voloch, M., Ladisch, M.R. and Tsao, G.T., 1982. Adsorption of ethanol-water mixtures
by biomass materials. Biotechnol. Bioeng., 24: 725-730.
Kovach, J.L., 1979. Gas-phase adsorption. In: P.A. Schweitzer (Editor), Handbook of Separation
Techniques for Chemical Engineers. McGraw-Hill, New York, 3.4-3.108.
Ladisch, M.R. and Dyck, K., 1979. Dehydration of ethanol: New approach gives positive energy
balance. Science, 205: 898-900.
Ladisch, M.R., Voloch, M., Hong, J., Bienkowski, P.R. and Tsao, G.T., 1984. Cornmeal adsorber
for dehydrating ethanol vapours. Ind. Eng. Chem. Process Des. Dev., 23 (3) : 437-443.
Levenspiel, O., 1972. Chemical Reaction Engineering. Wiley, New York, 501 pp.
Otten, L., 1985. Fuel ethanol dehydration. Report for Engineering and Statistical Research Insti-
tute, Contract File #34SZ.01843-1-EL09. Agriculture Canada, Ottawa, Ont., 105 pp.
Pan, C.Y. and Basmadjian, D., 1970. Analysis of adiabatic sorption of single solutes in fixed beds.
Pure thermal wave formation and its practical implications. Chem. Eng. Sci., 25 (11 ): 1653-1664.
Robertson, G.H., Doyle, L.R. and Pavlath, A.E., 1983. Intensive use ofbiomass feedstock in ethanol
conversion: the alcohol-water vapour phase separation. Biotechnol. Bioeng., 25: 3133-3148.
Turner, G.A., 1972. Heat and Concentration Waves. Academic Press, New York, 233 pp.
Van Genuchten, M.Th. and Alves, W.J., 1982. Analytical Solutions of the One-Dimensional Con-
vective-Dispersive Solute Transport Equation. USDA Agric. Res. Div., Tech. Bull. 1661, 151
pp.
Yang, C.M. and Tsao, G.T., 1982. Packed-bed adsorption theories and their applications to affin-
ity chromatography. In: A. Fiechter (Editor), Advances in Biochemical Engineering/
Biotechnology, 25. Springer, Berlin, 84 pp.

Potrebbero piacerti anche