Vous êtes sur la page 1sur 13

Improving bioethanol production from sugarcane: evaluation of distillation,

thermal integration and cogeneration systems


Marina O.S. Dias
a, b,
*
, Marcelo Modesto
c
, Adriano V. Ensinas
c
, Silvia A. Nebra
d
, Rubens Maciel Filho
a
,
Carlos E.V. Rossell
a, b
a
School of Chemical Engineering, University of Campinas, P.O. Box 6066, 13083-970 Campinas, SP, Brazil
b
Laboratrio Nacional de Cincia e Tecnologia do Bioetanol (CTBE), P.O. Box 6170, 13083-970 Campinas, SP, Brazil
c
Centre of Engineering, Modeling and Applied Social Sciences, CECS, Federal University of ABC, Rua Santa Adlia, 166, 09210-170 Santo Andr, SP, Brazil
d
Interdisciplinary Centre of Energy Planning - NIPE, University of Campinas, P.O. Box 6192, 13400-970 Campinas, SP, Brazil
a r t i c l e i n f o
Article history:
Received 1 March 2010
Received in revised form
9 September 2010
Accepted 10 September 2010
Available online xxx
Keywords:
Bioethanol
Process integration
Sugarcane
Exergetic analysis cost
Power generation
a b s t r a c t
Demand for bioethanol has grown considerably over the last years. Even though Brazil has been producing
ethanol fromsugarcane on a large scale for decades, this industry is characterized by lowenergy efciency,
using a large fraction of the bagasse produced as fuel in the cogeneration system to supply the process
energy requirements. The possibility of selling surplus electricity tothe gridor using surplus bagasse as raw
material of other processes has motivatedinvestments onmore efcient cogenerationsystems and process
thermal integration. In this work simulations of an autonomous distillery were carried out, along with
utilities demand optimization using Pinch Analysis concepts. Different cogeneration systems were
analyzed: a traditional Rankine Cycle, with steam of high temperature and pressure (80 bar, 510

C) and
back pressure and condensing steamturbines conguration, and a BIGCC (Biomass Integrated Gasication
CombinedCycle), comprisedbya gas turbine set operatingwithbiomass gas producedina gasier that uses
sugarcane bagasse as raw material. Thermoeconomic analyses determining exergy-based costs of elec-
tricity and ethanol for both cases were carried out. The main objective is to show the impact that these
process improvements can produce in industrial systems, compared to the current situation.
2010 Elsevier Ltd. All rights reserved.
1. Introduction
Increased interest on alternative fuels has been observed in the
past few years, as a result of increasing energy demand and fore-
casted depletion of fossil resources [1,2]. Global warming and the
consequent need to diminish greenhouse gases emissions have
encouraged the use of fuels produced from biomass [3], which is
the only renewable carbon source that can be efciently converted
into solid, liquid or gaseous fuels [4].
Bioethanol is presently the most abundant biofuel for automobile
transportation [5]. It is produced from fermentation of sugars
obtained from biomass, either in the form of sucrose, starch or
lignocellulose. Sugarcane is so far the most efcient rawmaterial for
bioethanol production: the consumption of fossil energy during
sugarcane processing is much smaller than that of corn [6]. One of
the main by-products generated during sugarcane processing is
sugarcanebagasse, whichis usually burnt inboilers for productionof
steamand electrical energy, providing the energy necessary to fulll
the process requirements. The use of efcient technologies for
cogeneration coupled with optimization of bioethanol production
process allows the production of surplus bagasse [7], which may be
used as a fuel source for electricity generation or as raw material for
producing bioethanol and other biobased products [8,9]. Different
cogeneration systems, such as the Rankine Cycle with condensing
steam turbines and those based on gasication technologies, may
improve the amount of electricity producedfromsugarcane bagasse,
thus allowing its sell to the grid.
Optimizationof bioethanol productionprocess canreduce energy
consumption, consequently increasing sugarcane bagasse avail-
ability. Thermal integration using Pinch Technology can reduce hot
andcoldutilities requirements, thus optimizingenergyconsumption
of the bioethanol production process [5].
Separation is the step where major costs are generated in process
industry [10]. In the case of bioethanol production, distillation
process may be optimized to reduce steam consumption on column
reboilers, therefore reducing production costs. The distillation
process commonly employed in Brazilian distilleries is based on the
same conguration used for decades, on which atmospheric pres-
sures are adopted; in a new conguration presented in this work,
* Corresponding author. Laboratrio Nacional de Cincia e Tecnologia do Bio-
etanol (CTBE), P.O. Box 6170, 13083-970 Campinas, SP, Brazil. Fax: 55 19 3518 3164.
E-mail address: marina.dias@bioetanol.org.br (M.O.S. Dias).
Contents lists available at ScienceDirect
Energy
j ournal homepage: www. el sevi er. com/ l ocat e/ energy
0360-5442/$ e see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.energy.2010.09.024
Energy xxx (2010) 1e13
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
wine is fed to distillation columns operating under vacuum pres-
sures, where ethanol-rich streams containing from 40 to 50 wt%
ethanol are produced; these streams are fedtoatmospheric-pressure
rectication columns, where hydrous ethanol (approximately 93 wt
%) is produced. Reduction of steam consumption on distillation
column reboilers is then achieved by the use of a double-effect
distillation system, since different temperature levels are obtained in
column condensers and reboilers, allowing the thermal integration
of these equipments [11].
In order to evaluate the consequences of introducing a double-
effect distillation system in ethanol production from sugarcane,
simulations of bioethanol production processes were carried out
using software UniSim Design fromHoneywell. Steamdemand was
evaluated for each case (bioethanol production using conventional
or double-effect distillation columns), after thermal integration
using Pinch Analysis was carried out. The results obtained were
used to assess two different cogeneration systems: Rankine Cycle
operating with steamat high temperature and pressure (480

C and
80 bar) and a BIGCC, on which sugarcane bagasse is gasied. Even
though the gasication technology is not yet commercially feasible,
the possibility of improving electricity production in this case
should be evaluated. Simulations of the cogeneration systems were
carried out using Gate Cycle and EES (Engineering Equation Solver).
Thermoeconomic analyses and determination of exergy-based
costs for ethanol and electricity were assessed for both cases.
2. Bioethanol production process
In this work an autonomous distillery, on which all sugarcane
processed is used to produce ethanol, for the production of one
million liters of anhydrous ethanol per day was considered. This is
a typical industrial scale unit, where around 12,000 ton of sugar-
cane (TC) per day are crushed, producing around 85 L of anhydrous
ethanol per ton of sugarcane.
Simulations were carried out using the commercial software
UniSim Design from Honeywell.
2.1. Chemical and physical properties
In order to represent sugarcane composition (displayed in
Table 1) more accurately, some hypothetic components that are not
part of UniSim database were created as described in a previous
work [12]: bagasse components (cellulose, hemicellulose and
lignin); sand, with properties considered equal to those of SiO
2
;
impurities, represented by potassium salts and aconitic acid; input
materials such as phosphoric acid and lime; calciumephosphate,
the mainsalt formedduringlimingoperation, of great importance in
removal of impurities during juice settlement; minerals, repre-
sented by K
2
O [13]; and yeast. Properties for these hypothetic
components were obtained in Wooley and Putsche [14] and Perry
and Green [15]. All reducing sugars are considered dextrose; all
other components (water, sucrose, ethanol, carbondioxide, glycerol,
succinic acid, acetic acid, isoamyl alcohol, sulfuric acid, MEG
(monoethyleneglycol)) are part of UniSim database.
NRTL (Nom-Random Two Liquid) was the model chosen for
calculating activitycoefcient of the liquidphase, andthe equationof
state SRK (SoaveeRedlicheKwong) for the vapor model. It was
veried that the NRTL model was the one that calculated elevation of
the boiling point of sucrose solutions with larger accuracy, when
compared to UNIQUAC (Universal Quasi Chemical) or Pen-
geRobinson equation of state, as represented in Fig. 1. For the
extractive distillation process, employed in ethanol dehydration, the
model UNIQUAC and equation of state SRK were used.
2.2. Main aspects of the production process
Production of anhydrous bioethanol from sugarcane is
comprised by the following main steps: reception and cleaning of
sugarcane, extraction of sugars, juice treatment, concentration and
sterilization, fermentation, distillation and dehydration. All these
steps were represented on the simulation [12]. Some process
improvements were considered in this work, such as: for sugarcane
cleaning, use of a dry-cleaning system, insteadof one basedonwater
use, what decreases sugar losses as well as water consumption;
efcient juice treatment, considering addition of phosphoric acid
and lime, what increases juice purity and consequently improves
the fermentation stage; concentration of juice on multiple effect
evaporators, decreasing process steam consumption; decreased
fermentation temperature (28

C), which allows the production of
wine with higher ethanol content, consequently diminishing sugar
and ethanol losses, as well as vinasse generation; a double-effect
distillation process, which allows thermal integration between
columns condensers and reboilers; optimization of the extractive
distillation process with MEGfor anhydrous bioethanol production,
considering feed of hydrous ethanol on the vapor phase (in the
industry, hydrous bioethanol is obtainedina condensedphase inthe
conventional distillation process, followed by vaporization prior to
the extractive distillation process, what leads to increased and
unnecessary steam consumption). A block-ow diagram of the
bioethanol production process is depicted in Fig. 2. Description of
the different steps of the bioethanol production process from
sugarcane is presented in the following subsections.
2.3. Sugarcane reception and cleaning, extraction of sugars and
juice treatment
A dry-cleaning system is used to clean sugarcane received in the
factory, removing about 70% of dirt before it is fed to the mills.
Extraction of sugars is done using mills, where sugarcane juice and
Table 1
Composition of sugarcane received in the factory.
Component Content (wt%)
Sucrose 13.30
Fibers 11.92
Reducing sugars 0.62
Minerals 0.20
Impurities 1.79
Water 71.57
Sand 0.60
0 10 20 30 40 50 60 70 80 90
100
102
104
106
108
110
Experimental
NRTL
Peng Robinson
UNIQUAC
)
C

(
t
n
i
o
P
g
n
i
l
i
o
B
Mass fraction of sucrose (%)
Fig. 1. Boiling point of sucrose solutions: experimental data [16] and values calculated
using the process simulator with NRTL, PengeRobinson and UNIQUAC.
M.O.S. Dias et al. / Energy xxx (2010) 1e13 2
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
bagasse are obtained. Water is used to improve sugars recovery, in
a process termed imbibition [13]. A physical treatment is used to
remove sand and ber from the juice, in which screens and hydro-
cyclones are used [13]. Juice then receives a chemical treatment
consisting of addition of phosphoric acid, lime, heating and settle-
ment to remove other impurities. In the settler, mud and claried
juice are obtained. A lter is used to recover some of the sugars
contained in the mud, and the ltrate is recycled to the process prior
to the second heating operation.
2.4. Juice concentration and sterilization
Claried juice contains around 15 wt% solids, so it must be
concentrated before fermentation in order for the product to contain
an adequate ethanol content that allows reduction of energy
consumption during purication steps. Concentration is done on
a 5-stage multiple effect evaporators (MEE) up to 65 wt% sucrose.
Vapor purges may be done if needed, in order to supply heat to other
operations. Only part of the juice is concentrated, and nal juice
contains around 22 wt% sucrose. In order to avoid contamination in
fermentation, juice is sterilized and cooled till fermentation
temperature (28

C).
2.5. Fermentation
Simulations were based on the Melle-Boinot (feed-batch with
cell recycle) process. Sterilized juice is added to the fermentors
along with yeast stream. In the fermentor sucrose is inverted to
glucose and fructose, which are consumed by the yeast producing
ethanol, CO
2
and other products, such as higher alcohols, organic
acids, glycerol and yeast. Fermentation gases are collected and
washed in an absorber for ethanol recovery, while wine containing
yeast cells are centrifuged to recover yeasts. The yeast milk stream
receives an acid treatment prior to be fed back to the fermentor, in
order to avoid contamination. Wine is mixed with alcoholic
solutions from the absorber before being fed to the distillation
columns.
Fermentation was carried at 28

C, in order to obtain wine with
higher ethanol content (around 10 wt%), decreasing ethanol and
sugar losses and energy consumption during the purication step
[17]. An alternative cooling method (absorption with lithium-
ebromide) was considered to supply cool water to maintain this
low fermentation temperature.
2.6. Distillation and dehydration
The conventional distillation process consists of a distillation
column comprised by columns A, A1 and D, and a rectication
column comprised by columns B and B1. Wine obtained in fermen-
tation is pre-heated and fed to column A1, which is located above
column A and below column D, as shown in Fig. 3. Pressure on the
distillation columns range from 133.8 to 152.5 kPa, and on the
rectication columns from 116 to 135.7 kPa.
In the distillation columns ethanol-rich streams containing
around 40 wt% ethanol (phlegms) are obtained, as well as vinasse
and 2nd grade ethanol. In the rectication column hydrous ethanol
Fig. 2. Block-ow diagram of the bioethanol production process from sugarcane in an
autonomous distillery.
A1
Vapour
Phlegm
Vinasse
R
R
Top
D
Liquid
phlegm
R
R
R
D
Top D
recycle
R
Gases
2nd grade
ethanol
A
B,
B1
Hydrous
Ethanol
Phlegmasse
Fusel
oil
Warm
wine
Fig. 3. Conventional distillation process conguration.
M.O.S. Dias et al. / Energy xxx (2010) 1e13 3
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
and residues like phlegmasse and fusel oil are obtained. Hydrous
ethanol in vapor phase is produced on top of column B.
Since water and ethanol form an azeotrope with concentration
of 95.6 wt% ethanol at 1 atm, an extractive separation process with
MEG was used to produce anhydrous bioethanol. In this process
two columns are used: an extractive column, where hydrous
ethanol and a suitable solvent is used, anhydrous bioethanol is
produced on the top and a solvent solution on the bottom, and
a recovery column, in which extractive solvent is recovered.
Extractive column operates at atmospheric pressure (101.325 kPa),
while the recovery column operates at 20 kPa, in order to avoid
high temperature and solvent decomposition. The solvent is cooled
before fed to the extractive distillation. Since the bottom temper-
atures of both extractive and recovery columns are relatively high
(136 and 150

C, respectively), both reboilers need to operate with
high pressure (6 bar) steam.
A1
Vapour
phlegm
Bottom
A
R
R
Liquid
phlegm
R
R
R
R
Warm
wine
Hydrous
ethanol.
Fusel
oil
Phlegmasse
B,
B1
Extractive
Recovery
MEG
Anhydrous
Ethanol
Water
Solvent
R
D
Gas
A
Boilup-A
Vinasse
R
2nd
grade
ethanol
Fig. 4. Double-effect distillation e integration of column A reboiler to column B and extractive column condensers.
Table 2
Initial and nal temperature, heat ow of streams considered for thermal integration.
Streams Conventional distillation Double-effect distillation
Hot streams T
initial
(

C)
T
nal
(

C)
Heat ow
(kW)
T
initial
(

C)
T
nal
(

C)
Heat ow
(kW)
Sterilized juice 130 28 27,606 130 28 27,602
Fermented wine 28 24 5440 28 24 5063
Vinasse 112 35 27,426 65 35 10,547
Anhydrous ethanol cooling 78 35 9070 78 35 8983
Vapor condensates 107 50 10,499 109 50 12,418
Condenser column B 82 82 24,297 82 82 23,656
Condenser extractive column 78 78 9224 78 78 13,734
Condenser column D 81 30 662 41 26 4482
Condenser recovery column 82 64 290 78 63 295
Cold streams T
initial
(

C) T
nal
(

C) Heat ow (kW) T
initial
(

C) T
nal
(

C) Heat ow
(kW)
Raw juice 30 70 21,062 30 70 21,096
Limed juice 76 105 20,535 74 105 20,667
Juice for sterilization 96 130 9287 96 130 9309
Centrifuged wine 28 82 23,161 28 48 8636
Reboiler column A 112 112 38,553 65 65 37,389
Reboiler column B 108 108 6725 108 108 5248
Reboiler extractive column 111 137 8432 111 137 12,881
Reboiler recovery column 150 150 1704 150 150 1794
M.O.S. Dias et al. / Energy xxx (2010) 1e13 4
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
In the double-effect distillation system, the distillation columns
operate under vacuumpressures (19e25 kPa), while the rectication
columnoperates at atmospheric pressures (101.325e135.7kPa), what
makes temperature onthe bottomof column Areachabout 65

C and
temperature on the top of column B reach 78

C. Consequently,
integration of column B condenser and column A reboiler is possible,
and the rectication column condenser may work as the distillation
column reboiler, what provides a reductioninsteamconsumption on
distillation. Nevertheless, since phlegms are obtained in the vapor
phase in column A, they must be compressed before being fed to
column B, what is done by using steamcompressors. Since column A
reboiler duty is greater than that of column B condenser, this inte-
gration cannot supply all the heat necessary to the adequate opera-
tion of the distillation columns. Given that temperature on top of the
extractive column from the extractive distillation process reaches
78

C, it is possible to use anhydrous bioethanol vapors to provide the


remaining heat necessary to an adequate operation of column A. This
integration in the double-effect distillation is shown in Fig. 4.
3. Process integration analysis
The Pinch Point Method described by Linnhoff et al. [18] was used
to analyze the streams of the process which are available for thermal
integration. The method developed by works of Hohmann [19],
Umeda et al. [20] and Linnhoff and Flower [21,22] uses ent-
halpyetemperature diagrams torepresent theprocess streams andto
nd the thermal integration target for them, considering a minimum
approach difference of temperature (Dt
min
) for the heat exchange.
The thermal integration target indicates the minimum
requirements of external hot and cold utilities for the process. The
Fig. 5. Composite Curves for the conventional distillation case.
Fig. 6. Composite Curves for the double-effect distillation case.
M.O.S. Dias et al. / Energy xxx (2010) 1e13 5
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
analysis is useful to represent all the streams available for heat
exchange in one diagram, being possible to evaluate the effect of
modications in the process parameters before the heat exchange
network design.
Thermal integration analysis of the bioethanol production
process considered the hot and cold streams parameters deter-
mined in the simulation of the autonomous distillery for both
conventional and double-effect distillation. Thermal integration
between rectication and distillation columns has an important
impact on the steam balance, promoting changes in the integration
possibilities. Data of the streams available for thermal integration
are presented in Table 2.
Some streams with less than 1000 kW of heat ow were not
considered in the analysis because of their low thermal integration
potential. The evaporation system was considered as a separated
systemwhich is integrated with the background process in a second
step. This procedure previouslyapplied by Ensinas [23] in sugarcane
mills simplies the analysis, maximizing the vapor bleeding use and
reducing the total exhaust steam consumption in the process.
As it was previously mentioned, in this analysis a single-effect
lithium bromide absorption system, with COP 0.65, was considered
to provide the cold utility requirements of the fermentation step,
using for that vapor from the third effect of the evaporation system
as heating source to produce the necessary cold water.
Thus, following the procedure presented by Linnhoff et al. [18]
the CC (Composite Curve) (Figs. 5 and 6) could be drawn and the
targets determined. Fig. 7 shows the GCC (Grand Composite Curve)
of double-effect distillation case considering the integration of the
background with the evaporation system after the thermal inte-
gration. The Pinch Point temperature in this case was found at
113

C, considering a global Dt
min
of 10

C for all the streams of the
background process, and a Dt
min
of 4

C for the evaporation vapor
streams. The software presented by Elsevier [24] was used for
calculation of the targets and drawing of CC and GCC curves.
4. Cogeneration systems
Traditionally, cogeneration systems employed in Brazilian sugar-
cane mills are based on the Rankine Cycle [25]. Sugarcane bagasse is
used as a fuel to supply thermal, mechanical andelectrical demandof
the sugar and ethanol production process. These plants used to
operate withsteamat lowlevels of pressure and temperature (20 bar
and 350

C) and back pressure steam turbines, resulting in low
energetic efciency, high bagasse consumption, low bagasse surplus
and a small electricity surplus, or even none. After the Brazilian
electrical crisis in 2001, independent energy producers were allowed
to sell electricity to the grid, thus providing the sugarcane industry
the possibility of improving electricity generation from biomass. As
a consequence, newprojects of sugarcane mills considered the use of
Rankine Cycles with steam at higher levels of temperature and
pressure, condensing steam turbines and the use of all the bagasse
generated in the mills as a fuel to produce electricity, selling the
surplus electricity to the grid. In the following sections a description
of the cogeneration systems studied in this work is made.
Fig. 7. Grand Composite Curves for the double-effect distillation case (background/
foreground process integration).
Fig. 8. Conguration of the cogeneration system based on Rankine Cycle: back pressure steam turbines.
M.O.S. Dias et al. / Energy xxx (2010) 1e13 6
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
4.1. Traditional cogeneration system in sugarcane mills (Rankine
Cycle)
Figs. 8 and 9 showtwo basic congurations of the Rankine Cycle
used in cogeneration systems traditionally employed in Brazil. The
rst conguration (Fig. 8) displays a boiler, a back pressure steam
turbine, deaerator and pumps; the second conguration (Fig. 9)
shows the same set of equipment; however, a condensing steam
turbine is employed, therefore a condenser is added to system. The
cogeneration systemprovides steam(2.5 and 6 bar of pressure) and
electricity to supply the anhydrous ethanol production process. The
ethanol production process receives sugarcane (stream11 on Fig. 8)
producing anhydrous ethanol (stream 12), vinasse (stream 13) and
sugarcane bagasse (stream 14), and requires 6 bar (stream 5) and
2.5 bar steam (stream 3) besides electrical energy (We). The
bagasse produced in the mills has a moisture content of 50% and is
used to supply energy to the cogeneration system(stream10). 7% of
the bagasse is separated to provide energy for start-ups (stream15)
and the remaining fraction is available for other uses (stream 16).
Sugarcane bagasse ow was calculated using Eq. (1).
m
bagasse

x
w
m
cane
(1)
where: x is the percentage of ber in sugarcane; w is the moisture
content of the bagasse. The values used for ber content of sugar-
cane and moisture of sugarcane bagasse are 12% and 50%,
respectively.
Nowadays, bagasse surplus is used to increase the production of
electricity, mainly in condensing systems as shown in Fig. 9.
However, when production of ethanol fromlignocellulosic materials
(second generation ethanol) becomes feasible, bagasse surplus may
be used to increase ethanol production from sugarcane, using the
same cultivatedarea throughpretreatment andhydrolysis processes.
The use of a Rankine Cycle was assessed through a thermoeco-
nomic analysis using software EES

; the main parameters used in


the simulation are shown in Table 3.
4.2. Cogeneration systems based on BIGCC (Biomass Integrated
Gasication Combined Cycle) in sugarcane mills
The use of the BIGCC in sugarcane mills has been investigated
for the past 15 years [26e33]. Other works assess the combined use
of biomass gas and natural gas in cogeneration plants, in order to
overcome the problems found in the BIGCC applied in sugarcane
mills [34e39]. The use of gas turbines with biomass is an option as
well [40e46].
A conguration of the BIGCC system employed in sugarcane
mills is illustrated in Fig. 10. The system consists of a gasier,
a dryer, a gas turbine set, an HRSG(Heat Recovery SteamGenerator)
and back pressure steam turbines that provide steam to supply the
demands of the process at both 2.5 and 6 bar. Similarly to the
previous case, sugarcane bagasse is produced during anhydrous
ethanol production (stream10 on Fig. 10) and fed to an atmospheric
gasier, on which biomass gas production takes place. The biomass
gas that leaves the gasier (stream 15) is compressed in a biomass
gas compressor and fed to a combustion chamber, on which it
reacts with compressed air (stream 2) producing stack gases
(stream 3) and supplying power for compressor and generator and,
consequently, producing electric energy. Stack gases from the gas
turbine (stream 4) are fed in the HRSG producing steam at 480

C
and 80 bar, which supplies the thermal demand of the ethanol
production process through steam turbines that produce low
pressure steam (stream 17, which consists of 6 bar steam and
stream 19, 2.5 bar steam). The stack gases that leave the HRSG
(stream 5) are used to dry the bagasse before it is fed in the gasier.
One of the most important steps of the simulation of BIGCC is
the modeling of the gasier. In order to simulate the behavior of the
BIGCC two important parameters must be determined: lower
heating value of biomass gas and adequate moisture content of the
bagasse after the drier. In order to obtain these parameters,
a computational tool kindly provided by Pellegrini and Oliveira Jr.
[47] is employed. The model is based on a study carried out by Fock
and Thomsem [48]. The input variables are sugarcane bagasse
composition (47.7% carbon, 5.8% hydrogen and 46.5% oxygen) [49]
Fig. 9. Conguration of the cogeneration system based on Rankine Cycle: condensing steam turbines.
Table 3
Main parameters adopted in the simulation of the Rankine Cycle.
Parameters Value
Sugarcane processed (ton/h) 493
Anhydrous ethanol production (L/TC) 85
Vinasse production (L/L of anhydrous ethanol) 11
Lower heating value of bagasse (kJ/kg) 7500
Isentropic efciency of pumps 0.85
Isentropic efciency of steam turbines 0.80
First law efciency of boiler 0.85
Electric power demand in the process (kWh/TC) 28
Fraction of bagasse for start-ups (%) 7
Efciency of electrical generators 0.98
M.O.S. Dias et al. / Energy xxx (2010) 1e13 7
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
and moisture. Gasication was considered to occur in the presence
of atmospheric air. In fact, the best gasication conditions are
present when a mixture of oxygen and steam is employed;
however, costs for separating oxygen from atmospheric air for use
in power generation are prohibitive.
The compositionof biomass gas as a functionof sugarcanebagasse
moisture is representedinFig. 11. The products of bagasse gasication
are CO, CO
2,
H
2
O, CH
4
, N
2
and H
2
. Molar fraction of CH
4
is considered
constant and that of the other components vary with bagasse mois-
ture. Decrease of bagasse moisture from50to40%leads toanincrease
of the H
2
fraction; lower moisture values lead to a decrease of the H
2
fraction. Adifferent behavior is observedfor the molar fractionof both
CO and N
2
, which increase when moisture content decreases. Molar
fractions of H
2
Oand CO
2
decrease along withmoisture; this behavior
affects the lower heating value of biomass gas as well as the gasi-
cation temperature, as shown in Fig. 12.
Proles of the gasication temperature and lower heating value
of the biomass gas as a function of bagasse moisture are depicted in
Fig. 12. Increase of the CO molar fraction and decrease of water
molar fraction gives rise to an increase on the lower heating value
of biomass gas. In addition, if the temperature is too low, there will
not be enough energy to start up the gasication process. According
to Reed and Gaur [50], the gasication process requires a minimal
temperature of approximately 800

C. Thus, the minimal value of
sugarcane bagasse moisture is 25%, as illustrated in Fig. 12.
Sugarcane bagasse produced during sugarcane processing has
a moisture content of 50%. Therefore, in order to be employed in
Fig. 10. Conguration of the cogeneration system based on the BIGCC cycle.
Fig. 11. Biomass gas composition as a function of bagasse moisture. Fig. 12. Lower heating value and biomass gas temperature versus bagasse moisture.
M.O.S. Dias et al. / Energy xxx (2010) 1e13 8
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
a gasication cycle, the moisture content of bagasse must be
reduced to 25%. The bagasse drying process takes place at the drier
(equipment IV depicted in Fig. 10). The drier makes use of the stack
gases from the HRSG (stream 5), and its main parameters are
temperature and composition of the stack gases and the outlet
minimal temperature of the gases leaving the device (which is
assumed to be equal to the stack gases adiabatic saturated
temperature). The energy balance in the drier is shown in Eq. (2)
m
5
h
5
m
10
h
10
m
6
h
6
m
14
h
14
0 (2)
where: m
10
m
bd
m
w10
, m
bd
represents the dry bagasse owand
m
w10
the amount of water present inthe bagasse; m
14
m
bd
m
w14
,
m
w14
corresponds to the amount of water in the bagasse that leaves
the drier and m
6
m
5
m
w6
, on which m
w6
is the amount of water
removed from the bagasse.
It is convenient to dene the moisture of bagasse through
streams 10 (bagasse fed to the drier) and 14 (dried bagasse), as
shown in Eq. (3).
w
10

m
w10
m
10
and w
14

m
w14
m
14
(3)
The values adopted in this simulation consider the moisture
content of the stream 14 equal to 15% [47]. However, since the
temperature (75

C, considered constant) of stream 6 (stack gases
leaving the drier) depends on the composition of the stack gases,
and it is necessary to reach 15% moisture on stream 15 (bagasse
leaving the gasier), energy must be provided to reach the required
moisture level through a convenient value of temperature of the
stack gas from the HRSG (stream 5). Usually, the value used in the
HRSG is close to 150

C [51]. However, this value is too lowto assure


a convenient bagasse drying, as shown in Fig. 13. Increasing stack
gases temperature leads to suitable levels of moisture for gasica-
tion; however, decrease of this temperature implies reduction on
the capacity of steam production for the ethanol production
process. In order to assure bagasse moisture of 15%, the minimum
temperature of the stack gases must be 215

C for this simulation, as


depicted in Fig. 13.
Ina previous study, Modesto et al. [52] showedthree proposals to
use the BIGCC cycle in ethanol plants. The rst two proposals used
a commercial gas turbine (GE LM 2500) adapted to use biomass gas
as fuel; however, the thermodynamic conditions adopted could not
supply the steam demand (340 kg/TC) without using a supplemen-
tary fuel (natural gas) in the HRSG. The third proposal, which
considers a gas turbine designed for biomass gas, allowed the
elimination of the supplementary fuel; however, the operating
conditions of the gas turbine penalize the performance of electric
generation. According to Modesto et al. [52], a suitable value of
steam demand to make a BIGCC cycle feasible is 230 kg/TC. Thus,
thermal integration and the use of a double-effect distillation
system, as described in this work, will improve the feasibility of the
BIGCC cycle by not penalizing electric generation.
In order to execute the thermoeconomic analysis of the BIGCC
cycle the software Gate Cycle (mass and energy balances) and
EES (exergy and exergy cost balances) were used. The main
parameters used in the simulation of BIGCC cycle are shown in
Table 4. In this simulation, a gas turbine using only biomass gas is
simulated using the Gate Cycle

Software, developed by GePower.


Palmer et al. [42] provided some suggestions for the use of biomass
gas in the operation of this gas turbine. The main parameters
analyzed were pressure ratio and temperature at the outlet of the
combustion chamber. The biomass gas from gasier is compressed
in a compressor and enters a combustion chamber along with an air
owat the same pressure. Both react in a combustion chamber and
the gases produced expand in a power turbine producing electrical
energy. It is important to mention that the exclusive use of biomass
gas in the gas turbine needs a catalytic combustion chamber
designed to operate with fuels of low heating values. Proposals of
catalytic combustion chambers can be found at Witton et al. [53]
and Forzatti [54]. Another aspect that needs to be mentioned is
that this study does not consider the energetic consumption of an
important step of the gasication process, the cleaning of gases that
leave the gasier. The cleaning of stack gases is a very important
step to allow the use of these gases in a gas turbine, in order to
assure suitable levels of impurities to avoid corrosion and scorching
in the blades of the turbine.
5. Exergetic cost analysis
In order to assess the two cogeneration systems proposals, an
exergetic cost analysis was performed. This analysis employs mass,
energy, exergy and exergy costs balances in all components of the
plant. Thus, it is possible to determine energetic consumption,
irreversibility generation and the exergetic cost of the products,
through a convenient cost allocation method.
Eqs. (4)e(6) show mass, energy and exergy balances for
a generic control volume, respectively, according to Kotas [55].
X
_ m
in

X
_ m
out
0 (4)
_
Q
_
W
X
_ m
in
h
in

X
_ m
out
h
out
0 (5)
_
Q

1
T
T
o

_
W
X
_ m
in
e
in

X
_ m
out
e
out

_
I (6)
In order to determine steam, stack gases, air, biomass gas and
water exergy, Eq. (7) was used
0 5 1
0 7 1
0 9 1
0 1 2
0 3 2
0 9
0 1 1
0 3 1
50 45 40 35 30 25 20 15
S
t
a
c
k

g
a
s
e
s

t
e
m
p
e
r
a
t
u
r
e

f
r
o
m

H
R
S
G

(

C
)
) % ( e r u t s i o m e s s a g a B
Fig. 13. Bagasse moisture as a function of stack gases temperature from HRSG.
Table 4
Parameters used in simulations of BIGCC cycle.
Parameter Value
Production of biomass gas (kg/kg of bagasse) 2.23
Lower heating value (kJ/kg) 5100
Pressure ratio 18
Isentropic efciency of compressors 0.85
Isentropic efciency of power turbine 0.92
Isentropic efciency of steam turbine 0.9
Isentropic efciency of pumps 0.85
Temperature of gases leaving the HRSG (

C) 215
Temperature of gases leaving the drier (

C) 75
M.O.S. Dias et al. / Energy xxx (2010) 1e13 9
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
e h
i
h
o
T
o
s
i
s
o
e
0
ch
(7)
where: h
i
enthalpy at point i; h
o
enthalpy of reference;
s
i
entropy at point i, s
o
entropy of reference; e
ch
chemical
exergy of water. The values used for the temperature and pressure
of reference are 25

C and 1 bar, respectively. The chemical exergy
values are determined by Szargut et al. [56].
The exergy values of sugarcane bagasse were calculated by the
methodology described in Sosa-Arnao and Nebra [57]; exergy of
the ethanolewater mixture employed the methodology described
in Modesto et al. [58].
The methodology used to perform the exergetic cost analysis
is the Theory of Exergetic Cost proposed by Lozano and Valero
[59]. This methodology can be used to determine the exergetic
and monetary cost of each ow that compose the system. The
evaluation of the exergetic and monetary costs of a cogeneration
system in a sugar plant evaluating the inuence of the price of
the main fuel (sugarcane bagasse) in steam production and
electricity costs was presented by Sanchez and Nebra [60]. The
use of the thermoeconomic methodology to assess the exergetic
cost of sugar in the production process was described by Fer-
nandez Parra [49].
The exergetic cost calculation is made through cost balance
equations in each component, as shown by Eq. (8).
X
k
in
E
in

X
k
out
E
out
0 (8)
where: k denes the exergy-based cost and E the total exergy
ow. The subscript in and out indicate the ows that enter and
leave the control volume, respectively.
The application of Eq. (8) in all control volumes forms a linear
equations set, where the number of variables is greater than the
number of equations. In order to obtain a set with a unique solution
it is necessary to add some additional equations. Cerqueira and
Nebra [61] reported in a simple way the postulates of the meth-
odology to dene these additional equations.
In the case of anhydrous ethanol production process, the exer-
getic cost balance equation is written by Eq. (9) (for the Rankine
Cycle) and Eq. (10) (BIGCC):
_ m
11
e
11
k
11
_ m
3
e
3
k
3
_ m
5
e
5
k
5
W
e
k
e
_ m
12
e
12
k
12
_ m
13
e
13
k
13
_ m
14
e
14
k
14
0 9
_ m
11
e
11
k
11
_ m
17
e
17
k
17
_ m
19
e
19
k
19
W
e
k
e
_ m
7
e
7
k
7
_ m
12
e
12
k
12
_ m
13
e
13
k
13
_ m
14
e
14
k
14
0 10
The set of additional equations were included following
the considerations proposed by Lozano and Valero [59]. For the
exergy-based costs of the inputs (sugarcane) a unitary value is
assigned, and therefore
k
11
1 (10a)
In BIGCC, stack gases that leave the drier have the exergy-based
cost null
k
6
0 (11)
All the irreversibility generation in the turbines must be carried
out by the exergy-based cost of electric power, and consequently
the exergy-based costs of the steam entering and leaving these
turbines are considered equal. Therefore, Eq. (12) (Rankine Cycle)
and Eq. (13) (BIGCC) are obtained
k
1
k
5
k
2
k
17
k
18
(12)
k
3
k
4
and k
9
k
17
and k
18
k
19
(13)
In the splitters, where no irreversibility generation takes place,
ows entering and leaving the valves have the same exergy-based
cost. Thus, Eqs. (14) and (15) are obtained, for the Rankine Cycle and
BIGCC, respectively.
k
2
k
3
k
4
; k
4
k
10
k
15
k
16
(14)
k
8
k
9
k
18
(15)
In the BIGCC, the exergy-based cost of stack gases that enter and
leave the HRSG(streams 4 and 5, respectively), is the same. Thus, all
the irreversibility generated in the HRSG is carried by exergy-based
cost of steam produced by HRSG (stream 8)
k
4
k
5
(16)
In the anhydrous ethanol production process, the following
considerations were made:
Table 5
Hot utilities demand before and after thermal integration for the studied congurations (A: conventional distillation without thermal integration; B: double-effect distillation
without thermal integration; C: conventional distillation with thermal integration; D: double-effect distillation with thermal integration).
Operation Saturated steam consumption (kg/h)
Conguration A Conguration B Conguration C Conguration D
2.5 bar 6 bar 2.5 bar 6 bar 2.5 bar 6 bar 2.5 bar 6 bar
1st juice heating 34,632 e 34,687 e e e e e
2nd juice heating 33,765 e 33,981 e e e e e
Multi-effect evaporation 56,270 e 60,706 e 59,254 e 68,519 e
Juice sterilization e 15,843 e 15,916 4056 4711 5427 4729
Reboiler column A 63,390 e e e 63,390 e e e
Reboiler column B 11,058 e 8629 e 11,058 8629 e
Reboiler extractive column e 14,426 e 22,039 e 14,426 e 22,039
Reboiler recovery column e 2915 e 3069 e 2915 e 3069
Total 199,115 33,184 138,003 41,024 137,758 22,052 82,575 29,837
Table 6
Thermal and electric energy demand of anhydrous ethanol production process.
Conguration 2.5 bar steam
consumption
(kg/s)
6.0 bar steam
consumption
(kg/s)
Electric energy
consumption
(kW)
Steam
Demand
(kg/TC)
A 55.31 9.22 0 468
B 38.33 4.43 4328 310
C 38.26 6.12 0 322
D 22.94 8.29 4328 226
M.O.S. Dias et al. / Energy xxx (2010) 1e13 10
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
i) The exergy-based cost of the steam that enters the process is
the same of the condensate obtained after heat exchange:
For the Rankine Cycle:
k
3
k
5
k
6
(17)
For the BIGCC cycle:
k
17
k
18
k
7
(18)
ii) The exergy-based cost of bagasse and vinasse (streams 14 and
13, respectively) is the same as that of the sugarcane (stream
11) that enters the process. Thus, all the irreversibility
generation is carried by the exergy-based cost of the ethanol.
As a result:
k
12
k
14
k
11
(19)
Thus, with the set equations above, the number of equations is
equal to the number of variables. The system was solved using the
EES

software.
6. Results and discussions
6.1. Process utilities requirements
After integration of the background process and the distribution
of the vapor bleeding, the hot utilities requirements could be
determined for each case. Four different congurations for the
anhydrous ethanol production process steam demand were
analyzed, considering conventional and double-effect distillation
with and without thermal integration. Two levels of steampressure
were considered for hot utilities, derived fromextractions at 6.0 bar
and 2.5 bar of pressure in the steam turbine. Steam consumption of
the analyzed cases is shown in Table 5, which summarizes the nal
results of hot utilities demand for each case.
As can be seen in Table 5, the thermal integration promoted
important steam reductions in the anhydrous bioethanol produc-
tion. For the mills with conventional distillation systemscheme the
reduction of the steam consumption at 2.5 bar can reach 31% with
the integration and 34% for the 6.0 bar steam.
When double-effect distillation columns are used in the mills
the reduction of steamconsumptionwith the thermal integration is
40% for the 2.5 bar steam and 27% for the 6.0 bar steam.
6.2. Evaluation of the cogeneration systems
Information regarding steam consumption, gathered from Table
5, and electric energy demand from the anhydrous ethanol process
are shown in Table 6. Exergetic costs and summarized results of the
evaluation are showed in Table 7.
The exergetic cost analysis in the Rankine Cycle with back
pressure steam turbine shows that the decrease of steam demand,
through the use of thermal integration and double-effect distilla-
tion (conguration D), leads to an increase of bagasse surplus;
however, lower electricity surplus is produced, mainly due to the
increase on electricity consumption of compressors in the double-
effect distillation system as well as to decreased steam consump-
tion. In fact, this conguration is the one that makes the best use of
bagasse for other applications, such as bagasse hydrolysis for
second generation bioethanol production: the exergy-based cost of
process steamand electricity remains constant, while exergy-based
cost of ethanol decreases due to the lower steam demand of the
process.
The case on which the Rankine Cycle is employed along with
condensing steam turbines shows higher surplus electricity than
the previous system. The possibility to use all the bagasse available
to generate electricity allows the production of an electricity
surplus of 79.61 kWh/TC, which corresponds to an increase of about
27%. However, the exergy-based costs are different than those of
the conventional cogeneration system used in Brazil: for electricity,
they are 5% higher, while for ethanol they are 0.5% lower.
The exergetic cost analysis of the BIGCC cycle was performed
only on conguration D, since this conguration has the lowest
steam demand (226 kg/TC), a value lower than the minimal
(230 kg/TC) foreseen by Modesto et al. [52]. The main difference of
the BIGCC cycle is the expressive electricity surplus produced in this
conguration. The value of 144.3 kWh/TC is 130% higher than
traditional Rankine Cycle with condensing steam turbines. In
addition, the exergy-based cost of electricity and ethanol is lower
(56 and 19%, respectively) than those of the traditional congura-
tion. Thus, the use of the BIGCC cycle leads to a higher electricity
surplus with low production exergy-based costs of electricity and
ethanol. Composition of costs of ethanol is found in Modesto et al.
[62]. All results are shown in Table 7.
7. Conclusions
A study of the bioethanol process modeling and thermal inte-
gration was described in detail in this paper. The use of the
commercial software UniSim Design as a modeling tool proved to
be useful for analysis of the energy integration opportunities,
particularly considering different distillation systems.
The thermal integration of the rectication and distillation
columns promoted important energy savings in the bioethanol
process as a whole. The analysis of the integrated plant with
a double-effect distillation system showed that this alternative
presents a lower process steam demand, when compared with
traditional distillation schemes. On the other hand, it requires
Table 7
Results of different congurations of cogeneration systems (STBP: Rankine Cycle with back pressure steam turbine; STCOND: Rankine Cycle with condensing steam turbine;
BIGCC: Biomass Integrated Gasication Combined Cycle).
Cogeneration System Process conguration Bagasse surplus (%) Electricity surplus (kWh/TC) Exergy-based cost
Steam Electricity Ethanol
STBP A 5.00 62.66 3.300 3.823 2.018
STBP B 37.05 23.86 3.300 3.824 1.861
STBP C 34.65 34.53 3.300 3.823 1.875
STBP D 54.03 6.23 3.300 3.817 1.790
STCOND A 0 62.62 3.402 3.841 2.238
STCOND B 0 71.74 3.536 3.966 2.093
STCOND C 0 78.8 3.525 3.956 2.017
STCOND D 0 79.61 3.611 4.036 2.007
BIGCC D 0 144.3 1.833 1.676 1.806
M.O.S. Dias et al. / Energy xxx (2010) 1e13 11
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
a more complex control system and higher investments, but this
analysis was not included in this paper. Another important issue
regarding this scheme is the high electric energy consumption of
the ethanol vapor re-compression step.
A comparison of different cogeneration systems for production
of steam and electricity in anhydrous ethanol production plants in
Brazil was also carried out. Three congurations of cogeneration
systems were assessed: traditional Rankine Cycles with back
pressure and condensing steam turbines and the BIGCC.
It was shown that the use of thermal integration techniques
allows a decrease of process steam demand and consequently
produces more surplus bagasse (in the case of back pressure steam
turbine) or increases electricity surplus (for the case where
a condensing steam turbine is considered), in both cases using the
traditional technology for cogeneration systems (Rankine Cycle).
The main differential of this study is that the feasibility of the
BIGCC in anhydrous ethanol production plants was investigated. A
previous study [52] showed that the BIGCC is feasible if the process
steam demand reaches about 230 kg/TC (the traditional value in
Brazilian plants is in the range of 380e450 kg/TC). Thermal inte-
gration and the use of a double-effect distillation system lead to
a signicant reduction of steam demand, reaching values below
230 kg/TC, thus indicating that the BIGCC cycle is feasible and may
provide a considerable increase on electricity surplus of the ethanol
plant.
This work assesses the feasibility of the use of BIGCC in anhy-
drous ethanol production process. It is true that gasication and gas
turbine technologies developed specically to use biomass gas are
not yet available commercially. However, this study shows that
from a thermal and exergetic analysis point of view, the use of this
technology allows an expressive increase of electricity generation
from sugarcane biomass.
Acknowledgments
The authors would like to thank CNPq, FAPESP and FINEP for
nancial support, Dr. Juan Harold Sosa-Arnao for valuable infor-
mation about modeling of the bagasse dryer and the reviewers for
their comments.
References
[1] Saxena RC, Adhikari DK, Goyal HB. Biomass-based energy fuel through
biochemical routes: a review. Renewable and Sustainable Energy Reviews
2009;13:167e78.
[2] Quintero JA, Montoya MI, Snchez OJ, Giraldo OH, Cardona CA. Fuel ethanol
production from sugarcane and corn: comparative analysis for a Colombian
case. Energy 2008;33:385e99.
[3] Gupta R, Sharma KK, Kuhad RC. Separate hydrolysis and fermentation (SHF) of
Prosopis juliora, a woody substrate, for the production of cellulosic ethanol by
Saccharomyces cerevisiae and Pichia stipitis-NCIM 3498. Bioresource Tech-
nology 2009;100:1214e20.
[4] Balat M, Balat H, z C. Progress in bioethanol processing. Progress in Energy
and Combustion Science 2008;34:551e73.
[5] Franceschin G, Zamboni A, Bezzo F, Bertucco A. Ethanol from corn: a technical
and economical assessment based on different scenarios. Chemical Engi-
neering Research and Design 2008;86:488e98.
[6] Macedo IC, Seabra JEA, Silva JEAR. Green house gases emissions in the
production and use of ethanol from sugarcane in Brazil: the 2005/2006 aver-
ages and a prediction for 2020. Biomass and Bioenergy 2008;32(7):582e95.
[7] Ensinas AV, Nebra SA, Lozano MA, Serra LM. Analysis of process steam
demand reduction and electricity generation in sugar and ethanol production
from sugarcane. Energy Conversion and Management 2007;48:2978e87.
[8] Buddadee B, Wirojanagud W, Watts DJ, Pitakaso R. The development of multi-
objective optimization model for excess bagasse utilization: a case study for
Thailand. Environmental Impact Assessment Review 2008;28:380e91.
[9] Walter A, Ensinas AV. Combined production of second-generation biofuels
and electricity from sugarcane residues. Energy 2010;35:874e9.
[10] Cardona CA, Snchez J. Fuel ethanol production: process design trends and
integration opportunities. Bioresource Technology 2007;98:2415e57.
[11] Pellegrini LF, Oliveira Jnior S, Burbano JC. Supercritical steam cycles and
biomass integrated gasication combined cycles for sugarcane mills. Energy
2010;35:1172e80.
[12] Dias MOS, Ensinas AV, Nebra SA, Maciel Filho R, Rossell CEV, Wolf Maciel MR.
Production of bioethanol and other bio-based materials from sugarcane
bagasse: integration to conventional bioethanol production process. Chemical
Engineering Research and Design 2009;87:1206e16.
[13] Chen JCP, Chou CC. Cane sugar handbook: a manual for cane sugar manu-
facturers and their chemists. John Wiley and Sons Inc.; 1993.
[14] Wooley RJ, Putsche V. Development of an ASPEN PLUS physical property
database for biofuels components. Report No. NREL/MP-425-20685. Golden,
Colorado: National Renewable Energy Laboratory. See also:, <http://www.
p2pays.org/ref/22/21210.pdf>; 1996.
[15] Perry RH, Green DW. Perrys chemical engineering handbook. 7th ed. New
York: McGraw-Hill; 1999.
[16] Hugot E. Handbook of cane sugar engineering. Amsterdam: Elsevier; 1972.
[17] Dias MOS. Simulation of ethanol production processes from sugarcane and
sugarcane bagasse, aiming process integration and maximization of energy
and bagasse surplus, [Simulao do processo de produo de etanol a partir
do acar e do bagao, visando a integrao do processo e a maximizao
da produo de energia e excedentes de bagao]. MSc dissertation: School
of Chemical Engineering, State University of Campinas; 2008 [in
Portuguese].
[18] Linnhoff B, Townsend DW, Boland D, Hewitt GF, Thomas BEA, Guy AR, et al.
User guide on process integration for the efcient use of energy. Warks,
England: The Institution of Chemical Engineers; 1982.
[19] Hohmann EC. Optimum networks for heat exchange. PhD thesis: University of
Southern California; 1971.
[20] Umeda T, Harada T, Shiroko K. A thermodynamic approach to the synthesis of
heat integration systems in chemical processes. Computers and Chemical
Engineering 1979;3:273e82.
[21] Linnhoff B, Flower JR. Synthesis of heat exchanger networks: I. Systematic
generation of energy optimal networks. AIChE Journal 1978a;24:633e42.
[22] Linnhoff B, Flower JR. Synthesis of heat exchanger networks: II. Evolutionary
generation of networks with various criteria of optimality. AIChE Journal
1978b;24:642e54.
[23] Ensinas AV. Thermal integration and thermoeconomic optimization applied to
industrial process of sugar and ethanol from sugarcane [Integrao trmica e
otimizao termoecmica aplicadas ao processo industrial de produo de
acar e etanol a partir da cana-de-acar]. PhD Thesis: School of Mechanical
Engineering, State University of Campinas; 2008 [in Portuguese].
[24] Elsevier Ltd. Pinch analysis spreadsheet. See also, http://books.elsevier.com/
companions/0750682604; 2007.
[25] Macedo I, Nogueira LFH. Assess of expansion ethanol production on Brazil, In:
Bio fuels e strategic issue of Brazilian government; 2005 [in Portuguese].
[26] Ogden JM, Hochgreb S, Hylton M. Steam economy and cogeneration in cane
sugar factories. International Sugar Journal 1990;92:131e40.
[27] Larson ED, Williams RH, Ogden JM, Hylton MG. Biomassegasier steam-injec-
ted gas turbine cogeneration for the cane sugar industry. In: Energy from
biomass andwastes XIV. Chicago: Institute of Gas Technology; 1991. p. 781e95.
[28] Walter ACS, Overend RP. Analysis of BIGeGT cycles in the sugarcane industry.
In: 10th biomass European congress, vol. 1. Wurzburg; 1998a. p. 1158e61.
[29] Walter ACS, Overend RP. Financial and environmental incentives: impact on
the potential of BIG-CC technology at the sugarcane industry, In: World
Renewable Energy Congress, vol. 3. Florence, Italy; 1998b. p. 1996e9.
[30] Larson ED, Williams RH, Leal MRLV. A review of biomass integrated-gasier/
gas turbine combined cycle technology and its application in sugarcane
industries, with an analysis for Cuba. Energy for Sustainable Development
2001;5(1):54e75.
[31] Turn SQ, Bain RL, Kinoshita CM. Biomass gasication for combined heat and
power in the cane sugar industry. International Sugar Journal
2002;104:268e73.
[32] Hassuani SJ, Leal MRLV, Macedo IC, editors. Biomass power generation:
sugarcane bagasse and trash. Piracicaba: United Nations Development Pro-
gramme (UNDP), Centro de Tecnologia Canavieira (CTC); 2005.
[33] Pellegrini LF, Oliveira Jr S. Exergy analysis of sugarcane bagasse gasication.
Energy 2007a;32(4):314e27.
[34] Rodrigues M, Walter A, Faaij A. Co-ring of natural gas and biomass gas in
biomass integrated gasication/combined cycle systems. Energy
2003a;28:1115e31.
[35] Rodrigues M, Walter A, Faaij A. Performance evaluation of atmospheric
biomass integrated gasier combined cycle systems under different strategies
for the use of low caloric gases. Energy Conversion and Management
2007;48(4):1289e301.
[36] Rodrigues M, Walter A, Faaij A. Techno-economic analysis of co-red biomass
integrated gasication/combined cycle systems with inclusion of economies
of scale. Energy 2003b;28:1229e58.
[37] Zamboni LM, Pellegrini LF, Tribess A, Oliveira Jr S. Comparative evaluation of
natural gas and sugarcane bagasse based cogeneration systems. In: The 18th
international conference on efciency, costs, optimization, simulation and
environmental impact of energy systems, vol. 3. Trondheim: Tapir Academic
Press; 2005. p. 1105e12.
[38] Zanetti AA, Pellegrini LF, Oliveira Jr S. Thermoeconomic analysis of a BIGCC
cogeneration system using natural gas and sugarcane bagasse as comple-
mentary fuels. In: The 20th international conference on efciency, costs,
M.O.S. Dias et al. / Energy xxx (2010) 1e13 12
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024
optimization, simulation and environmental impact of energy systems, vol. 1.
Padova: Sevizi Graci Editoriali; 2007. p. 829e38.
[39] Walter A, Llagostera J, Oliveira Jr S, Pellegrini LF. Gas turbine integration to
sugarcane trash boilers: comparative energy analysis. In: The 20th interna-
tional conference on efciency, costs, optimization, simulation and environ-
mental impact of energy systems, vol. 1. Padova: Sevizi Graci Editoriali;
2007. p. 801e8.
[40] Walter AC, Llagostera J. Exergetic and thermoeconomic analysis of
biomass integrated gasier/gas turbine power cycles. In: Proceeding of
the ASME COGEN-turbo power conference. Vienna, Austria; Aug 23e25
1995.
[41] Walter A, Llagostera J. Off-design operation of cored combined cycles. In:
Proceedings ECOS 2003 e the 16th international conference on efciency,
costs, optimization, simulation and environmental impact of energy systems.
Copenhagen, Denmark; June 30e2 July 2003. p. 1443e50.
[42] Palmer CA, Erbes MR, Pechtl PA. Gate cycle performance analysis of the
LM2500 gas turbine utilizing low heating values. In: Proceedings of IGTI ASME
Cogen-Turbo; 1993.
[43] Kapat JS, Agrawal AK, Yang T. Air extraction in a gas turbine for integrated
gasication combined cycle (IGCC): experiments and analysis. Journal of
Engineering for Gas Turbines and Power 1997;119(1):20e6.
[44] Korobitsyn MA, Jellema P, Hirs GG. Possibilities for gas turbine and waste
incinerator integration. Energy 1999;24:783e93.
[45] Ferreira SB, Nascimento MAR, Pilidis P. The use of biomass fuels in gas turbine
combined cycles: gasication vs externally red cycle. In: III world climate &
energy event, vol. 1. Rio de Janeiro, Brazil; 2003. p. 1e8.
[46] Nascimento MAR, Ferreira SB, Pilidis P. A comparison of different gas turbine
concepts using biomass fuel. New Orleans: The American Society of
Mechanical Engineers; 2001.
[47] Pellegrini LF, Oliveira Jr. S. Exergy analysis of sugarcane bagasse gasication.
In: Proceedings of ECOS 2005-the 18th international conference on efciency,
costs, optimization, simulation and environmental impact of energy systems.
Trondheim, Norway; June 20e22 2005. p. 393e400.
[48] Fock F, Thomsen KBP. Modelling a biomass gasication system by means of
EES. Kgs. Lyngby, Denmark: The Scandinavian Simulation Society, Technical
University of Denmark; 2000.
[49] Fernandez Parra MI. Exergoeconomic methodology in sugar process [Meto-
dologia de analise exergoeconomica do processo de fabricao de acar]. PhD
Thesis: School of Mechanical Engineering, State University of Campinas; 2003
[in Portuguese].
[50] Reed TB, Gaur S. A survey of biomass gasication 2000. The National
Renewable Laboratory and The Biomass Energy Foundation; 2001.
[51] Kehlhofer R, Bachmann R, Nielsen H, Warner J. Combined cycle gas &
steam turbine power plant. Tulsa, Oklahoma: PennWell Publhishing
Company; 1999.
[52] Modesto M, Nascimento MAR, Nebra SA. Proposals for BIGCC cycle utilization in
an ethanol production plant from sugarcane. In: ECOS 2007 e the 20th inter-
national conference on efciency, costs, optimization, simulation and envi-
ronmental impact of energy systems. Padova, Italy; 25e28 June 2007. p. 761e8.
[53] Witton JJ, Noordally E, Przybylski JM. Clean catalytic combustion of low heat
value fuels from gasication process. Chemical Engineering Journal 2003;91
(2e3):115e21.
[54] Forzatti P. Status and perspectives of catalytic combustion for gas turbine.
Catalysis Today 2003;83(1e4):3e18.
[55] Kotas TJ. The exergy method of thermal plant analysis. Malabar Florida:
Krieger Publishing Cobarny; 1995.
[56] Szargut J, Morris DR, Steward FR. Exergy analysis of thermal, chemical and
metallurgical process. New York, USA: Hemisphere Publishing Co.; 1988.
[57] Sosa-Arnao JH, Nebra SA. Exergy of sugarcane bagasse. In: Proceedings of 14th
European biomass conference & exhibition. biomass for energy, industry and
climate protection, 17e21 October 2005. Paris, France; 1995.
[58] Modesto M, Nebra SA, Zemp RJ. A proposal to calculate the exergy of non ideal
mixtures ethanolewater using properties of excess. In: Proceedings of the
14th European conference and exhibition: biomass for energy, industry and
climate protection. Paris, France; 17e21 October 2005.
[59] Lozano MA, Valero A. Theory of the exergetic cost. Energy 1993;18
(9):939e60.
[60] Sanchez MG, Nebra SA. Thermoeconomic analysis of a cogeneration system of
a sugar mill plant. In: Proceeding of ECOS 2002 e 15th international confer-
ence on efciency, costs, optimization, simulation and environmental impact
of energy systems, vol. I. Berlin, Germany; 2002, p. 258e7.
[61] Cerqueira SAAG, Nebra SA. Cost attribution methodologies in cogeneration
systems. Energy Conversion and Management 1999;40(15e16):1587e97.
[62] Modesto M, Zemp RJ, Nebra SA. Ethanol production from sugarcane: assess of
possibilities of decrease of thermal energy consumption through exergetic
cost analysis. Heat Transfer Engineering 2009;30(4):272e81.
M.O.S. Dias et al. / Energy xxx (2010) 1e13 13
Please cite this article in press as: Dias MOS, et al., Improving bioethanol production from sugarcane: evaluation of distillation, thermal inte-
gration and cogeneration systems, Energy (2010), doi:10.1016/j.energy.2010.09.024

Vous aimerez peut-être aussi