Vous êtes sur la page 1sur 7

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/267224712

SIMULATION OF DISTILLATION PROCESS FOR BIOETHANOL PRODUCTION


CONSIDERING FERMENTATION BY-PRODUCTS

Article

CITATIONS READS

0 487

3 authors:

Tassia L. Junqueira Rubens Maciel Filho


Laboratório Nacional de Ciência e Tecnologia do Bioetanol University of Campinas
49 PUBLICATIONS   871 CITATIONS    620 PUBLICATIONS   6,168 CITATIONS   

SEE PROFILE SEE PROFILE

Maria Regina Wolf Maciel


University of Campinas
215 PUBLICATIONS   1,917 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Lactic Acid Production from Lignocellulosic Hydrolysate Using Cell Recycling Fermentation of Bacillus coagulans View project

Polymerization in situ of Nylon 6 View project

All content following this page was uploaded by Rubens Maciel Filho on 16 April 2015.

The user has requested enhancement of the downloaded file.


SIMULATION OF DISTILLATION PROCESS FOR BIOETHANOL PRODUCTION CONSIDERING
FERMENTATION BY-PRODUCTS

Tassia L. Junqueira, Rubens Maciel Filho, Maria R. W. Maciel

School of Chemical Engineering, State University of Campinas, Brazil

Abstract: Bioethanol represents an important alternative source of energy, since it reduces


polluting gases emissions, especially when compared to fossil fuels. As a mean to increase its
economic efficiency as fuel, optimization proposals are essential, mainly on ethanol
concentration and dehydration, which usually require a significant amount of energy.
Bearing this in mind, the focus of this work is separation of a mixture proceeding from
fermentation stage. Usually the published works deal with ethanol and water mixture but in
practice this is not the case when industrial units for ethanol production are taken into
account. This mixture effluent from the fermentator, after centrifugation, is sent to
conventional distillation columns in order to concentrate ethanol near the azeotropic value.
However, the great number of components, like ethanol, water, glycerol, isoamyl alcohol,
glucose, acetaldehyde and furfural, and the strong interaction between them make the
column’s simulation more complex. The work was carried out in Aspen Plus simulator and a
thermodynamic models adequacy evaluation was made to represent suitably the separation
process units. Process optimization was realized in order to minimize energy consumption,
ethanol losses and achieving product specification. Further, results have shown that
fermentation by-products impact the general process behavior and energy consumption.

Keywords: Bioethanol, conventional distillation, by-products, optimization, thermodynamic


models

1. INTRODUCTION

Bioethanol (ethanol from biomass) is an important renewable fuel, since it contributes to the reduction of negative
environmental impacts generated by the fossil fuels utilization. However, only in the last years, with the increasing
environmental concerns and the periodic crisis in some of the larger oil exporting countries together with the
increase in the oil price, bioethanol has become a viable and realistic alternative in the energy market (Cardona and
Sánchez, 2007). To point up the present importance of ethanol, in 2007, Brazil’s production was 19 millions m³,
while United States produced almost 25 millions m³ (RFA, 2007).

Fuel ethanol is obtained from sugarcane in some tropical countries like Brazil and India, from beet molasses in some
European countries and the main feedstock in the USA is starch from corn. Bioethanol production from sugarcane is
a well known process. It fundamentally consists of: sugarcane cleaning, extraction of sugars, juice treatment
(removal of sand, fiber and impurities), concentration, sterilization, fermentation, distillation and dehydration.

In the distillation section, hydrated ethanol (93 wt%) is obtained and it can be used as fuel or be sent to dehydration
section in order to produce anhydrous ethanol (99.3 wt%), which is added to gasoline. This blend allows a best
oxidation of hydrocarbons, reducing the amounts of polluting gases released into the atmosphere. Nevertheless, to
achieve the anhydrous ethanol specification, the azeotrope (approximately 96 wt% of ethanol) must be eliminated.
Therefore, further operations are required and extractive and azeotropic distillations are possible and usual solutions.

Besides being an attractive and sustainable energy source, bioethanol is also raw material for chemical industries.
From ethanol it is possible to obtain high quality acetaldehyde, acetic acid, ethyl acetate, ethylene and from them a
huge amount of chemicals, including polymers, making the sugarcane very interesting from the environmental point
of view and an economically attractive raw material to be used to obtain chemicals (Rivera et al., 2008).

2. PROCESS SIMULATION

Taking into account the fact that distillation operations are responsible for the major energy consumption, the focus
of this work is the concentration of the mixture proceeding from the fermentation stage, named wine. This mixture
contains a lot of components, such as water, ethanol, glycerol, glucose, isoamyl alcohol, furfural and acetaldehyde.

In typical Brazilian’s sugar mills, traditional columns configuration consists of distillation and rectification stages.
The distillation phase takes place in three columns, in the top of each other: A, A1 and D. Here, ethanol is partially
extracted from the wine and the other remaining parts leave at the bottom as phlegm (between 45 and 55%, mass
basis) being directed fed to column B-B1 for further extraction (Meirelles, 2006). The volatile contaminants are
concentrated in column D and removed from its top. The effluent at the bottom of column A is called vinasse or
stillage. The rectification consists in the concentration of the phlegm to about 93% in the column B-B1, and the
removal of more impurities (higher alcohols, aldehydes, etc). The product of this phase is hydrated ethanol and the
effluent is the phlegmasse that has a high water concentration and ethanol traces.

This work was carried out in Aspen Plus simulator and simulations considered this traditional columns
configuration. Components mass flows and conditions of the wine were also based on industrial sugar mill’s data.
These data allowed the component mass fractions calculation. All wine specifications are shown in Table 1.

Table 1 Conditions, components mass flows and mass fractions of the wine

Component Mass (kg/h) Mass fraction


Water 190413 0.920
Ethanol 15155 0.073
Glycerol 814 0.004
Isoamyl alcohol 316 0.002
Glucose 146 0.001
Isobutanol 69.4 -
n-Propanol 21.7 -
Acetone 2.29 -
Acetaldehyde 1.83 -
Acetal 1.05 -
Acetic acid 0.626 -
n-Butanol 0.525 -
Total mass flow (t/h) 206.9
Temperature (°C) 81.00
Pressure (bar) 1.393

In spite of the great number of components, only a few components have a significant mass fraction: water, ethanol,
glycerol, isoamyl alcohol and glucose. These components were taken into account in the initial simulations. It is
important to emphasize that the wine composition is strongly dependent on fermentation conditions. Under high
temperatures or low pH, it is common to find out furfural and more acetaldehyde in the mixture. Then, new
simulations were performed considering such compounds in the mixture.

Since the feed composition was known, thermodynamic models adequacy to represent the system was evaluated.
This was done comparing the vapor-liquid equilibrium (VLE) results given by some models present in Aspen Plus
database and available experimental data (Gmehling and Onken, 1977). This study analyzed NRTL, UNIQUAC,
PSRK and UNIFAC and all the missing interaction parameters were estimated using UNIFAC. NRTL was the most
suitable model with a good agreement with the experimental data, while the other tested models do not allow such
good predictions.
Columns configuration was introduced in Aspen Plus simulator as shown in Figure 1. Two cases were simulated; the
first considered the composition given in Table 1, while in the second, acetaldehyde and furfural were taken into
consideration. Acetaldehyde and furfural mass fractions were assumed to be 0.002 and 0.005, respectively, so water
mass fraction resulted in 0.913. These values were also obtained from industrial plant. Other components mass
fractions were the same shown in Table 1. Both cases considered the configuration illustrated in Figure 1,
nevertheless some adaptations were necessary. The configuration process was adapted to be used in the first
simulation, since there were no components more volatile than ethanol in the wine. As a result, the stream
proceeding from the column D condenser was sent straight to the splitter to be divided into the streams ALCOHOL2
and REFLUX. Simulations were carried out looking for achieving mass fraction specifications and minimizing
ethanol losses.

Figure 1. Configuration of the simulated process.

In Table 2, columns conditions and specifications are summarized. Stages numbering is initiated in the top stage or
condenser, when it is present and coupled reboiler, as in column A, is considered the last stage.

In these simulations, FUSEL stream refers to fusel oil, that consists of heavy alcohols, such as amyl alcohol. The
fusel oil withdrawal position was determined by analysing amyl alcohol mass fraction in liquid phase. Column B-B1
composition profiles can be seen in Figures 2 and 3 for cases 1 and 2, respectively.

In Figures 2 and 3, two peaks are observed in stages 21 and 45. In the first case, both peaks have almost the same
amyl alcohol mass fraction, so sidestreams were withdrawal from these stages. In the second case, 21st stage peak
has not the highest mass fraction in the column, so only the sidestream from stage 45 was withdrawal.
Table 2 Distillation columns conditions and specifications
Column A
Number of stages 19
Feed stage 1
Top pressure (kPa) 139.3
Bottom pressure (kPa) 152.5
PHLEGM-V withdrawal stage 2 (vapor phase)
Distillate rate (kg/h) 11300
PHLEGM-V rate 30000
Column A1
Number of stages 8
Feed stage: WINE 1
TOP-A 8
Top pressure (kPa) 136.3
Bottom pressure (kPa) 139.3
Column D
Number of stages 6
Feed stage: REFLUX 1
TOP-A 6
Top pressure (kPa) 133.8
Bottom pressure (kPa) 136.3
Column B-B1
Number of stages 46
Feed stage: PHLEGM-L 22
PHLEGM-V 22
VAPOR 46
Top pressure (kPa) 116.0
Bottom pressure (kPa) 135.7
FUSEL withdrawal stages* 21 and 45 (liquid phase)
FUSEL rates (kg/h)* 500 and 300
Distillate rate (kg/h) 16100
* In the second case, FUSEL was only withdrawal in the stage 45.

Figure 2. Amyl alcohol mass fraction profile for case 1.


Figure 3. Amyl alcohol mass fraction profile for case 2.

3. COMPARISON BETWEEN SIMULATIONS

In order to compare both simulations, some objectives were defined. Product ethanol mass fraction had to be at least
0.93 and ethanol should not be found in phlegmasse and vinasse streams. Further, product rate was fixed to be
16100 kg/h in both cases. Finally, the main streams conditions and composition of both cases are shown in Table 3.

Table 3 Conditions and composition of the main streams

Variable Simulation 1 Simulation 2


Stream name PRODUCT PHLEGMASS VINASSE PRODUCT PHLEGMASS VINASSE
Temperature (°C) 81.7 105.6 111.9 81.5 104.9 111.8
Pressure (bar) 1.160 1.357 1.525 1.160 1.357 1.525
Mass Flow (kg/h) 16100 13110 176931 16100 13851 176277
Mass Fraction
Water 0.069 0.987 0.994 0.061 0.956 0.991
Glucose 0 0 0.001 0 0 0.001
Ethanol 0.931 0 0 0.930 0 0
Glycerol 0 0 0.005 0 0 0.005
Isoamyl alcohol 0 0.013 0 0 0.015 0
Furfural - - - 0 0.030 0.003
Acetaldehyde - - - 0.009 0 0

From Table 3 analysis, it can be inferred that acetaldehyde and furfural presence influenced the columns behavior.
The hydrous ethanol, in the second simulation, contains a considerable amount of acetaldehyde. The phlegmasse
stream has a lower water content in the second simulation, because of furfural presence. Nevertheless the vinasse
water content was not so influenced by the acetaldehyde and furfural addition. In order to complete this analysis, the
amounts of energy required in both cases were compared. These data are shown in Table 4.
Table 4. Amount of energy required in each equipment and total energy consumption
Energy required (Gcal/h) Difference
Equipment
Simulation 1 Simulation 2 (%)*
Reboiler (column A) 17.32 17.31 0.0564
Condenser (column B-B1) -10.18 -13.74 -35.05
Condenser (column D) -0.0036 -0.1035 -2789
Reboiler (column B-B1) 2.829 6.482 -129.2
Cooler - -0.0140 -
Total cooling energy -10.18 -13.86 -36.16
Total heating energy 20.14 23.79 -18.09
* The difference was taken into account the first simulation.

Comparing the amounts of energy required in both simulations, it follows that when furfural and acetaldehyde are
taken into account the process requires more energy, either for cooling or heating. Column A required energies are
almost equal in both cases and cooler is only necessary in the second case to condense acetaldehyde, which requires
a small amount of energy. Although the relative difference between the required energies of column D condenser is
enormous, the required energy is not high, so it does not influence the total difference. Since column B-B1 relative
differences are remarkable and the required energy is very significant, this column is responsible for the major part
of this difference.

3. CONCLUDING REMARKS

Binary xy diagrams showed that the binary interactions are quite complex, with azeotrope and formation of two
liquid phases. As a result, it is important to have a systematic study of the adequacy of the thermodynamic models to
represent conveniently the system.
Additionally, since other components, such as acetaldehyde and furfural, are rarely taken into account, the results in
this work allow to conclude that these components, even though in little amount, impact the general column
behavior including the energy consumption.
This study also showed that simulation including acetaldehyde and furfural is more complex, which means the
process simulation convergence is more sensitive to operation changes.

4. ACKNOWLEDGMENTS

The authors acknowledge FAPESP for financial support.

REFERENCES

Cardona, C. A. and Sánchez, O. J. (2007). Bioresource Technology 98, 2415-2457.


Gmehling, J., Onken, U. (1977) Vapor-Liquid Equilibrium Data Collection – DECHEMA Chemistry Data Series.
Meirelles, A. J. A. (2006). Ethanol Production Workshop, Engineering School of Lorena - USP, Lorena, Brazil.
RFA - Renewable Fuels Association (2007). Available from: www.ethanolrfa.org/industry/statistics. Accessed May
12, 2008.
Rivera, E. C., Costa, A. C., Lunelli, B. H., Wolf-Maciel, M. R. and Maciel Filho, R. (2008). Applied Biochemistry
and Biotechnology 148, 163-173

View publication stats

Vous aimerez peut-être aussi