Recuperación y Caracterización de Etanol Celulósico A Partir de La Fermentación de Bagazo de Caña de Azúcar

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Chemical Engineering Research and Design 196 (2023) 568–576

Available online at www.sciencedirect.com

Chemical Engineering Research and Design

journal homepage: www.elsevier.com/locate/cherd

Recovery and characterization of cellulosic ethanol


from fermentation of sugarcane bagasse
]]
]]]]]]
]]


Celina K. Yamakawa a, , Sebastian T. Rojas b, William E. Herrera c,
Carlos E.V. Rossell d, Maria Regina Wolf Maciel c, Rubens Maciel Filho c
a
Department of Biotechnology and Biomedicine, Technical University of Denmark, Søltofts Plads, Building 223,
2800 Kongens Lyngby, Denmark
b
Institute for Molecular Bio Science, Goethe University Frankfurt, Frankfurt, Germany
c
School of Chemical Engineering, State University of Campinas, Campinas, SP, Brazil
d
Interdisciplinary Center on Energy Planning, State University of Campinas, Campinas, SP, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: Industrial production of ethanol by fermentation using renewable feedstock such as su­
Received 19 December 2022 garcane stalks has been demonstrated as a sustainable fuel chain in Brazil. This work
Received in revised form 29 May focused on the production of cellulosic ethanol from sugarcane bagasse in a pilot scale
2023 unit by applying the current bioprocessing strategies with the aim of recovering and
Accepted 26 June 2023 characterizing the end products. The feedstock was pretreated at 190 °C and a residence
Available online 28 June 2023 time of 10 min. Enzymatic hydrolysis was performed with commercial cellulolytic en­
zymes. Fermentation’s substrates were formulated with hydrolysate and supplemented
Keywords: with 8%wt sugarcane molasses. The fermentations were set up to mimic the conventional
Sugarcane bagasse industrial fermentation in Brazil’s ethanol distilleries at high cell density with cell re­
Lignocellulosic biomass cycling. The fermentation resulted in a reproducible performance by the yield of 0.49 g/g,
Cellulosic ethanol productivity of 6.96 g/(L∙h), and cell viability of 95.3%. Ethanol was recovered in a lab-scale
Fermentation distillation batch system. Distilled fractions showed higher content of higher alcohols and
Cell recycling sulfur content than the standard specification of ANP (National Agency of Petroleum -
Distillation Brazilian Agency) for ethanol fuel. The distillation bottom product (vinasse) presented
Vinasse most characteristics suitable for fertilizer or biogas applications, except for sodium and
Ethanol fuel standard sulfate content. Therefore, for a successful technology, transference processing adjust­
ments should be made to make the product commercially suitable and the side stream
compatible for disposal as fertilizer or digestion for biogas production.
© 2023 Institution of Chemical Engineers. Published by Elsevier Ltd. All rights reserved.

produced during the fermentation of sugars by Saccharomyces


1. Introduction
cerevisiae yeast. Currently, the most used sources of sugars
are sugarcane, corn, and sugar beet (Zabed et al., 2017). Par­
Biofuels play a crucial role in the advancement of the dec­
ticularly, Brazilian bioethanol distilleries are energy self-
arbonization of the economy, mitigating climate change by
sufficient, where the steam and electricity are generated
using renewable resources for the production of fuels (Letti
from bagasse and straw in the thermoelectric plant installed
et al., 2022). A classic example of biofuel is ethanol, which is
together with the production site (de Souza Dias et al., 2015).
Ethanol has been used intensely as an additive, fuel or
blended with conventional gasoline (Mizik, 2021). Beyond

Corresponding author. fuels, ethanol is widely used in food, pharmaceuticals, cos­
E-mail addresses: celinayamakawa@gmail.com, metics, detergents, and medical care areas. Also, it is a raw
cikiya@dtu.dk (C.K. Yamakawa).
https://doi.org/10.1016/j.cherd.2023.06.053
0263-8762/© 2023 Institution of Chemical Engineers. Published by Elsevier Ltd. All rights reserved.
Chemical Engineering Research and Design 196 (2023) 568–576 569

material to generate several industrial products of interest generation of the bottom product is between 10 and 15 L of
such as ethylene, propylene, and butadiene (Rossi et al., vinasse/L of ethanol (Bergmann et al., 2018). Therefore,
2021). Brazil and USA are the largest producers responsible ethanol production from sugarcane in Brazil has a noticeably
for 75% of ethanol production worldwide (Bergmann et al., high efficiency and well-established technology. But in the
2018). Nowadays, lignocellulosic biomass such as sugarcane light of using lignocellulosic biomass, introducing a new
bagasse, wheat straw, and cultivated wood have been sugar stream could cause impacts during fermentation, dis­
exploited for the production of ethanol (Guigou et al., 2019; tillation, effluent disposal, and the quality of the end product.
Soccol et al., 2010; Townsend et al., 2017) being called as For instance, lignocellulosic hydrolysates are generated
second-generation ethanol. Sugarcane bagasse is generated using different approaches, such as diluted acid hydrolysis
after extracting sucrose by crushing the sugarcane stalks. and saccharification, that bring a new sugar stream with the
This fiber residue contains 50% of moisture (Dias et al., 2012). presence of inhibitors, salts, and proteins. This work ad­
Today there is a surplus of bagasse due to the advances in dressed the application of sugarcane bagasse hydrolysate for
process efficiency like energy optimization and modern and ethanol production using the current methodology to ana­
more efficient equipment. Therefore, there is room to utilize lyze the characteristics of the final product (ethanol) and the
the lignocellulosic sugarcane residues as a feedstock to in­ effluent (vinasse) in comparison to the Brazilian fuel stan­
crease the productivity of ethanol without land-use compe­ dard regulation of ANP (National Agency of Petroleum -
tition with food crops. Integration of the current ethanol Brazilian Agency).
process and the second-generation processes is a realistic
approach and strategic principle to minimize the capital in­ 2. Material and methods
vestment, optimize the feedstock logistic, and promote the
advancement of biorefineries by producing added-value 2.1. Hydrolysate preparation
chemicals (Dias et al., 2013; Klein et al., 2017).
The polymeric structure of cellulose of bagasse can be Sugarcane bagasse was obtained from the Granelli sugar mill
hydrolysate by the application of heat, chemicals, and cel­ (Charqueada, São Paulo, Brazil). First, mineral impurities
lulolytic enzymes to obtain a stream rich in glucose together were removed from the feedstock using a vibratory sieve
with pentoses, organic acids, furan derivatives, and phenolic Multideck® (MultiVibro, BR). Three pretreatment campaigns
compounds. In the integration concept, this new stream of were performed to generate cellulosic hydrolysate for fer­
sugars is combined with the current stream of fermentation mentation. The pretreatment was performed in a 350-L
substrate formulated with sugarcane juice and molasses. stirred reactor (Pope Scientific Inc., USA) at 190 °C and 10 min
The configuration of the conventional bioprocess in Brazil of residence time using 10% (w/v) of solid loading, and 0.5%
comprises high cell density and cell recycling (Lopes et al., (v/v) of sulfuric acid. The reactor’s vessel was made of cor­
2016; Walker and Basso, 2020). Most of the installations apply rosion-resistant nickel-chromium-molybdenum wrought
fed-batch mode with 4 h of feeding followed by 2 h or more to alloy (Hastelloy C276). Pretreated bagasse was recovered in a
complete sugars exhaustion. At the end of fermentation, the 140-L Nutsche filter (Pope Scientific Inc., USA) made in
cells are recovered by centrifugation, washed at acidic con­ Hastelloy C22. The solid fraction, rich in cellulose and lignin
dition of pH 2–3 adjusted with sulfuric acid, and reused in the (cellulignin), was transferred to a basket centrifuge VTC 400/
following new batch. The amount of cell suspension corre­ 200 ha (Ferrum, CH) with capacity of 12.8-L. Water was in­
sponds to 1/3 of the final volume, and the rest is dedicated for jected in the basket for washing the cellulignin until the
substrate feeding. The current ethanol titer at the end of outlet water reached stable value of pH and %brix. The cel­
fermentation is between 8% and 12% (v/v), reaching product lulignin generated by three pretreatment campaigns was
yield of 92%. Ethanol is recovered and concentrated by dis­ transferred to the 350-L stirred reactor (Pope Scientific Inc.,
tillation. Conventional distillation consists of a set of 3 dis­ USA) for the enzymatic hydrolysis procedure. That was per­
tillation columns and 2 rectification columns (Dias et al., formed at 50 °C, pH of 4.8, adjusted with 2 M H2SO4 and 2 M
2011; Rossell et al., 2016). In the distillation columns is ob­ NaOH, during 72 h, 10% (w/v) of solid loading, and 10 FPU/mL
tained ethanol-rich stream containing approximately 40% (v/ of cellulolytic enzymes Cellic® Ctec2 (Novozymes, DK). After
v) ethanol, and the bottom product stream is named vinasse. completing the hydrolysis, the liquid fraction (cellulosic hy­
Vinasse is practically free of ethanol, but rich in salts and drolysate) was recovered in a 140-L Nutsche filter (Pope
organic matter. Currently, vinasse can be converted into Scientific Inc., USA) and concentrated in a wiped film eva­
biogas (Volpi et al., 2022), but it is mostly used in the fertili­ porator (Pope Scientific Inc., USA) at 80 °C and 475 mbar. The
zation of sugarcane crops at a limited amount to avoid soil hydrolysate concentration was done until it reached a sugars
salinization (Rossell et al., 2016). The ethanol-rich stream in a concentration of 212 g/L that corresponded to the equivalent
vapor phase goes to the rectification columns to concentrate of total sugars, in term of reducing sugars, to obtain a final
near the binary ethanol-water azeotrope. Either a heater ethanol concentration of 7.5°GL (59.2 g/L) in a configuration
exchange or direct steam injection is used in the distillation of an industrial fed batch at 90% of product yield (Basso
column. Moreover, during the distillation, impurities are re­ et al., 2010).
moved. The distillation columns are also designed to remove
separately secondary alcohols and incondensable compo­ 2.2. Substrate preparation
nents (CO2 and SO2). The rectification columns are con­
figurated with condensers to remove the compounds named The obtained concentrated hydrolysate was conditioned as a
fusel oil and higher oil streams (Rossell et al., 2016). Fusel oil substrate for fermentation. The conditioning steps and the
is composed mainly of 3-Methyl-1-butanol (amyl alcohol) substrate formulation were done in the same manner as
and 2-Methyl-1-propanol (isobutyl alcohol) (Mayer et al., previous work (Yamakawa et al., 2016) to meet the nutrient
2015). Higher oil is composed of n-propanol, isobutanol, and requirement for the yeast’s maintenance. The conditioning
n-butanol. The current distillation yield is 99.5%, and the step, called clarification, removes particles and proteins. This
570 Chemical Engineering Research and Design 196 (2023) 568–576

industrial procedure benefits yeast cell recycling by avoiding same manner as previous work (Rivera et al., 2017) but using
obstruction of centrifuges nozzles and the fermentation itself a carbon source of a mixture of sugarcane juice and molasses
by minimizing the foaming formation due to the presence of as cited previously.
protein, and minimizing the cell yield. Clarification treat­
ment relies on the pH drop with phosphoric acid, followed by
2.4. Ethanol fermentation
pH arising with lime and particle aggregation by adding non-
ionic polymer. The clarification treatment of concentrated
Ten repeated fermentations in fed-batch were performed in
hydrolysate was performed in a stainless steel 19.5 L-Bior­
a 2-L Bioflo 115 bioreactor (New Brunswick, USA) at 33 °C,
eactor (New Brunswick Scientific, USA) with the addition of
200 rpm, without any gas sparging. The first fed-batch was
0.4 mL/L of phosphoric acid (85%) followed by titration with
done by adjusting the cell concentration at 70 g/L (dry basis)
78 g/L calcium oxide solution until pH of 6.4 and heating up
in sterile tap water. The final weight corresponded to 600 g
to 95 °C. Afterwards, the temperature was immediately de­
and was transferred to the bioreactor. The substrate for
creased to 25 °C, and 4 ppm of non-ionic polyacrylamide was
feeding corresponded to 1400 g at a concentration of 195 g/L
added. After 10 min, the hydrolysate was centrifuged in a
of sugars. After completing each fermentation, yeast cells
basket centrifuge VTC 400/200 ha (Ferrum, CH). The hydro­
were recovered by centrifugation in a centrifuge Avanti J-26
lysate was distributed in 2-L Erlenmeyer flasks for steriliza­
XP (Beckman Coulter, EUA) at 13,225 x g, 15 min, and 20 °C.
tion in an autoclave AV-100 Plus (Phoenix, BR) at 121 °C for
The supernatant was kept refrigerated for further distilla­
15 min. The fermentation substrate was formulated with
tion. The cells pellet was resuspended in sterile tap water,
sterile hydrolysate and 8%wt of sugarcane molasses. Su­
and its concentration was adjusted to 70 g/L of cells in a total
garcane molasses support the nutrient demand for the
of 600 g. Then, the cell suspension was transferred back to
yeast’s maintenance during fermentation with cell recycling
the bioreactor’s vessel to perform acidic yeast treatment
(Yamakawa et al., 2016).
before new fed-batch fermentation. The yeast treatment was
performed at 33 °C, 800 rpm, and 0.2 L/min of air, followed by
2.3. Microorganism, inoculum, and yeast propagation
pH adjustment at 2.5 with 2 M sulfuric acid. After 30 min of
acidic treatment, the air supplier was cut off, the stirrer was
An isolated strain of yeast Saccharomyces cerevisiae originally
adjusted at 200 rpm, and the substrate feeding started. These
obtained from Santa Adélia sugarcane mill (Jaboticabal, São
3 stages of fermentation are shown in Fig. 1.
Paulo, Brazil) was used in this work. The stock culture was
maintained in YPD (10 g/L yeast extract, 20 g/L peptone and
20 g/L dextrose) at - 80 °C with 30% (v/v) glycerol. A pre-cul­ 2.5. Distillation
ture was prepared by transferring the stock culture to a 500-
mL Erlenmeyer flask containing 100 mL of YPD broth fol­ All supernatants recovered after centrifugation of the fer­
lowed by incubation at 33 °C and 250 rpm. After 24 h, an mented broth were combined for ethanol recovery in a batch
aliquot of pre-culture was transferred to 100 mL of inoculum distillation system Autodest® 800AC (iFISCHER, DE) con­
medium in a 500-mL Erlenmeyer flask. Inoculum medium trolled by Indusoft Web studioTM v6.216. The distillation
was prepared with 5.0 g/L of (NH4)2HPO4 and 80 g/L of sugars. column system comprehended a column packed with Pro-
The source of sugars corresponded to a mixture of sugarcane Pak® of 6 mm in 316 SS. The column was equipped with a
juice and molasses in a proportion of 79%wt of sugarcane silvered vacuum mantle and a thermometer for temperature
juice and 21%wt in terms of reducing sugars (Rivera et al., control. The temperature set-point was adjusted to 78 °C. The
2017). The inoculum was incubated in the same conditions as reboiler was equipped with an external mantle and two
the pre-culture. After 12 h, cells were recovered in the high- temperature sensors. Temperature set-points were adjusted
performance centrifuge Avanti J-26 XP (Beckman Coulter, at 100 °C (top jacket) and 170 °C (bottom jacket). A cryostat
USA) with rotor JLA-16.250 at 5509 x g, 10 °C for 15 min. The with a capacity of 360 Kcal at 0 °C was used in the condenser.
pellet was diluted with sterile tap water and transferred to The reflux ratio of 3 was adopted. This batch distillation
the 7.5-L bioreactor (New Brunswick, USA) for yeast propa­ system is equivalent to a column of 15 theoretical plates
gation to obtain the biomass needed for fermentation at high (ASTM D 2892). The distillate streams were collected at 5 time
cell density. The yeast propagation was conducted in the points between 38 °C and 80 °C.

Fig. 1 – Photos of cellulosic ethanol fermentation with cell recycling in 2-L bioreactor in fed-batch with cell recycling during
different steps a) Step of yeast cell treatment for 0.5 h, followed immediately by b) step of substrate feeding distributed in 6 h,
and followed by final c) step of sugars depletion for 2 h.
Chemical Engineering Research and Design 196 (2023) 568–576 571

2.6. Analytical methods cell conversion yield (YX/S) was obtained as the ratio between
produced cells and total sugars consumed. Sugars corre­
2.6.1. Feedstock composition, fermentation substrate and sponded to the sum of glucose and equivalent reducing su­
metabolites gars (glucose and fructose) from sucrose. Substrate to
The feedstock was characterized according to NREL’s stan­ product conversion yield (YP/S) was obtained as the ratio be­
dard procedure (Hames et al., 2008; Sluiter et al., 2012, 2008). tween the produced ethanol and total sugars consumed. The
The composition of hydrolysate and the fermentation me­ volumetric productivity (Qp) was calculated by dividing the
tabolites were analyzed by high-performance liquid chro­ ethanol produced per final volume of the fermented medium
matography (HPLC). Sucrose, glucose, fructose, and xylose and the total fermentation time.
were determined in an Agilent Infinity 1260 with IR detector
at 50 °C, Aminex column HPX-87 P 300 mm × 7.8 mm at 60 °C 3. Results and discussion
and 0. 5 mL/min of ultrapure Milli-Q water as eluent phase.
Acetic acid, ethanol, formic acid, and glycerol were de­ 3.1. Feedstock characterization and hydrolysate
termined in a Dionex Ultimate 3000 with IR detector Shodex composition
RI-101, Aminex column HPX-87 H 300 mm × 7.8 mm at 50 °C
and 0.5 mL/min of 5 mM sulfuric acid as eluent phase. Fur­ The composition of the lignocellulosic biomass corresponded
fural and 5-HMF were determined in an Agilent Infinity 1260 to in wt%: cellulose, 38.57 ± 6.23; hemicellulose, 25.10 ± 4.09;
with UV detector at 274 nm, Acclaim 120 - C18 150 × 4.8 mm lignin, 28.30 ± 0.97; acetic acid, 2.84 ± 0.36; extractives,
at 25 °C and 0.8 mL/ min of acetonitrile with water (1:8) with 3.13 ± 0.08; ashes, 5.42 ± 0.11. This composition agrees with
8% of acetic acid as eluent phase. the reported values (Canilha et al., 2012; Silva et al., 2011), but
the ash content found in this work was much higher and
2.6.2. Fermentation volatile metabolites and characterization justified the need for impurities removal by vibratory sieves.
of ethanol These mineral impurities can damage equipment such as
The composition of distillate products was quantified by gas pretreatment bioreactor and nozzle centrifuges used for
chromatography (GC) in a GC-2010 Shimadzu equipment and yeast cell recovery. The moisture content of biomass corre­
following the European Standard EN 15721. The distilled sponded to 19.42 ± 0.46%. The composition of the con­
streams were characterized by color (ASTM D 1209); acidic centrated cellulosic hydrolysate corresponded to in g/L:
(ABNT NBR 16047) using a potentiometric titration KEM glucose, 212.29; xylose, 11.69; glycerol, 0.10; acetic acid, 0.72;
(Kyoto Electronics); electrical conductivity at 25 °C (ABNT HMF, 0.092; furfural; 0.002.
NBR 10547) using Orion Star conductivity meter; sodium
content (ABNT NBR 10422); permanganate test (ASTM D 3.2. Ethanol fermentation
1363); sulfur content (ASTM D 5453).
The substrate used for ethanol fermentation was composed
2.6.3. Vinasse characterization in g/L: sucrose, 27.80; glucose, 121.19; fructose, 10.85; xylose,
The bottom product (vinasse) was characterized according to 8.46; acetic acid, 0.53; lactic acid, 4.76; glycerol, 0.15, HMF,
the Standard Method for the Examination of Water and 0.10; furfural, 0.05. Sucrose and lactic acid came from su­
Wastewater (SMEWW) 22nd edition. garcane molasses; lactic acid bacteria are the most common
contaminants found in sugar and ethanol production site
2.6.4. Yeast concentration, viability, and budding (Lino et al., 2020). Industrial yeast strain of Saccharomyces
The cell concentration in dry basis of the initial and final cerevisiae can well ferment sucrose, glucose, and fructose into
samples of each fermentation was determined gravime­ ethanol under anaerobic environment. Xylose will end as
trically in triplicate after centrifuging 1 mL of samples, residual sugar contributing to the organic matter in the dis­
washing the cell pellets twice with deionized water, and tillation bottom product. The organic acids and furan deri­
drying it at 80 °C until a constant weight was obtained in an vatives compounds inhibit yeast growth and affect cell
analytical scale. Yeast cell viability and budding were quan­ physiology significantly, such as decreasing intracellular pH
tified by classical staining technique with methylene blue and viability (Palmqvist and Hahn-Hägerdal, 2000). However,
(Lee et al., 1981) and counted using a Neubauer chamber in the inhibitory effects can be minimized in a fed-batch fer­
an optical microscope Eclipse CI-S (Nikon, JP). mentation where the microorganism will face a low con­
centration of inhibitors. Thus, this primary characterization
2.6.5. Fermentation monitoring of the substrate formulated with lignocellulosic biomass
During the course of fermentation, sugars and ethanol were hydrolysate can give an essential overview to follow the
monitored in real-time, soon after the sampling, in a mid- coming results of ethanol fermentation with cell recycling.
infrared spectrometer (Bruker Optics Inc., USA) that had been The fermentation profiles of all cycles were combined in a
calibrated accordingly (Rivera et al., 2017). Cell concentration single graphic as presented in Fig. 2. Cell concentration de­
was monitored by optical density at 600 nm in a spectro­ creased over time due to substrate feeding, whereas ethanol
photometer using water as a blank. Also, total soluble solids accumulated resulting in a final ethanol titer of
in %brix were measured using an optical refractometer 59.09 ± 2.39 g/L. The initial ethanol value of 17.89 ± 4.07 g/L
(Atago, JP). corresponded to the ethanol with the cells; ethanol was re­
leased after acidic cell washing. After this step, the substrate
2.6.6. Fermentation performance parameters was fed immediately over the washing water. Decreasing
The parameters of fermentation performance were calcu­ external pH plays an important role in the yeast cell per­
lated based on the weight of the transferred mass of the cell meability (Jones and Greenfield, 1987) that changes the in­
suspension and substrate to the bioreactor, and the mass of tracellular pH (Charoenbhakdi et al., 2016). Regarding the
fermented broth at the end of each fed-batch. Substrate to sugars, they accumulated up to 12 g/L until 6 h, which was
572 Chemical Engineering Research and Design 196 (2023) 568–576

Table 2 – Fractions of distilled product at different


column heating temperatures and sampling time-point
at 72–80 °C with corresponding volume and ethanol titer.
Fraction Temp. Volume Ethanol titer
(°C) (mL) (%v/v)

1 25–38 110 33.45


2 38–72 524 81.78
3 72–80 440 79.50
4 72–80 500 43.90
5 72–80 120 13.06

lost as residual sugars. Nowadays, on-line measurement


such as mid-infrared spectroscopy combined with real-time
analysis has been applied in several ethanol plants to elevate
the production management for making more accurate de­
cisions (Ethanol Producer Magazine, 2021).
Fig. 2 – Ethanol fermentation profile in fed-batch mode with
Therefore, the downscaling process of ethanol fermenta­
cell recycling: sugars accumulation (black square symbol
tion showed that the operation of ethanol plant using lig­
and line), cells dilution (orange circle symbol and line), and
nocellulosic biomass hydrolysate will deliver similar key
accumulation of produced ethanol (green triangle symbol
parameter indicators. In addition, 8%wt of sugarcane mo­
and line).
lasses were sufficient for cell maintenance and low growth.

kept constant during the last hour to complete the fermen­


tation. The residual sugars corresponded to xylose. More­ 3.3. Ethanol recovery and characterization
over, Table 1 shows the fermentation performance in terms
of cell and ethanol yields, ethanol titer, volumetric product The fermented media free of cells from all performed fer­
productivity, and cells viability and budding. The fermenta­ mentations totalized 15.3 L that was submitted to distillation
tion conditions benefited the product yield by reaching in a batch system. Five fractions were collected from the
0.49 ± 0.02 g/g in contrast to the cell yield of 0.04 ± 0.01 g/g. initial heating until stabilized at 80 °C. The volume of each
Ethanol titer corresponded to 59.09 ± 2.39 g/L or approxi­ fraction and its ethanol titer are shown in Table 2. The
mately to 7.5°GL as expected. Industrial yields have been equivalent amount in absolute ethanol corresponded to
reported to be 90% of ethanol efficiency and cell yield be­ 0.94 L. Dividing it by the lignocellulosic biomass gives 12.5 L
tween 0.04 and 0.05 g of cell/g of ethanol (Della-Bianca et al., ethanol/tonne of bagasse (dry basis), which is far from the
2013; Yamakawa et al., 2017). In this work, yields were higher expected value of 150 L ethanol/tonne of bagasse in the sce­
than typical industrial values, which could be attributed to nario of 90% efficiency for enzymatic hydrolysis (Yamakawa,
the better homogeneous system in a small-scale bioreactor. 2016). In this work, very low enzymatic hydrolysis conversion
In consequence, the volumetric productivity of ethanol was obtained due to the ineffective mixture of solids and
reached a value of 6.96 ± 0.21 g/(L·h), which is higher than enzymes in a conventional stirrer reactor. Customized re­
the typical industrial value of 2.6 g/(L·h) (Yamakawa et al., actors and operation strategies have been developed to
2019) for fed-batch operation. In addition, on-line measure­ maximize enzymatic hydrolysis of lignocellulosic biomass,
ment of sugars and ethanol led to more accurate data to for instance, membrane reactors and novel continuous mode
decide the end of the fermentation where all sugars were operation (Andrić et al., 2010; Lischeske and Stickel, 2019;
exhausted (Rivera et al., 2013; Yamakawa et al., 2019). It Narendranath, 2015; Olivieri et al., 2021). About the fractions
shortens the fermentation time and minimizes the sugars collected during the distillation, Fig. 3 shows the appearance

Table 1 – Fermentation performance parameters of 10 repeated fed-batch cycles with cell recycling in term of cell yield
(YX/S), product yield (YP/S), final ethanol titer, volumetric productivity (QP), percentage of cell viability, and percentage of
cell budding.
Cycle YX/S YP/S Ethanol titer QP Cell viability Cell budding
(g/g) (g/g) (g/L) (g/ (L·h)) (%) (%)

1 0.09 0.47 53.55 6.74 96.61 8.27


2 0.06 0.43 56.74 6.63 98.84 5.66
3 0.07 0.48 59.37 6.88 98.16 2.40
4 0.05 0.51 60.33 6.97 98.82 6.42
5 0.01 0.51 60.04 7.03 93.83 7.56
6 0.04 0.51 61.93 7.36 90.58 8.31
7 0.05 0.50 60.00 7.10 96.00 8.85
8 0.02 0.49 60.89 7.15 90.62 11.38
9 0.01 0.49 58.65 6.89 94.17 6.46
10 0.01 0.47 59.37 6.91 94.20 6.28
Mean 0.04 0.49 59.09 6.96 95.18 7.16
Deviation 0.01 0.02 2.38 0.21 3.06 2.36
Chemical Engineering Research and Design 196 (2023) 568–576 573

alcohols can be removed from the rectifier column to prevent


operational difficulties and an inability to produce high
quality ethanol (Rossell et al., 2016).
Characterization of the distilled fractions was done for
Fraction 3 and Fraction 4, as shown in Table 4, following the
standard specification of ANP (National Agency of Petroleum
- Brazilian Agency) for ethanol fuel. Fraction 3 presented
most of the desired characteristics such as total acidic,
electrical conductivity, and pH. The high color index can be
attributed to the hydrolysate itself that contained furfural, 5-
Fig. 3 – Photos of distilled fractions obtained at different HMF, and phenolic compounds, molasses, and byproducts
heating column temperatures and time-point when formed during the pretreatment step, also the sterilization in
reached 72–80 °C, left to right, fraction 1 (25 – 38 °C), 2 autoclave contribute to the formation of color due to de­
(38–72 °C), 3 (72–80 °C), 4 (72–80 °C), and 5 (72–80 °C). gradation of xylose at 121 °C. Clarification treatment applied
on the hydrolysate was not so effective for hydrolysate,
meaning that is necessary to adjust the current method for
Table 3 – Content of volatile compounds in the distilled hydrolysate-based substrate. Clarification process is com­
fractions.
monly applied in the sugar and ethanol processing for re­
Compound Fraction 2 Fraction Fraction Fraction 5 moving colloids particles and soluble proteins. Moreover, the
(mg/kg) 3 (mg/kg) 4 (mg/kg) (mg/kg)
composition of the hydroalcoholic fraction obtained in this
Butanol 65 34.4 3.8 < 10 work was slightly different from previous work that applied
3-Methyl-1- 1086 41.3 5.0 21 sugarcane juice fermentation using the same distillation
butanol system. The reported values (Alvarez et al., 2013) were:
Propanol 772 500 95 28
ethanol, 80% (w/w); 3-Methyl-1-butanol, 1.5% (w/w); acet­
2-Methyl-1- 1960 602 36.6 38
aldehyde, 22.5 mg/L; acetone, 7 mg/L; ethyl acetate, 30 mg/L;
propanol
Acetaldehyde 425 60 31 < 50 propanol, 70 mg/L; 2-Methyl-1-propanol, 325 mg/L; butanol,
Acetal 42 < 10 12.2 < 16 6 mg/L. The content of butanol and propanol of Fraction 3
Acetone 26 < 10 < 10 < 10 and Fraction 4 were pronounced higher compared to this
Ethyl acetate 207 10.5 < 10 < 10 work. Concerning other parameters to assess the quality of
ethanol, sulfur content did not meet the standard specifica­
tion of ANP. Sulfur content is not permitted for ethanol fuel.
of all fractions. Fraction 1 exhibited a whitish appearance Sulfur content originates in fermentation at the stage of
while Fraction 2 presented a yellowish appearance. These acidic cell washing with sulfuric acid. However, in this work,
colors can be attributed to the presence of impurities. an extra addition of sulfuric acid occurred during the titra­
Whereas Fractions 3, 4, and 5 were colorless. tion of enzymatic hydrolysis, and the sulfuric acid added
Regarding the volatile compounds besides ethanol, apart during the pretreatment was removed from the solids after
from Fraction 1, the superior alcohols were quantified intense washing with water. Therefore, adjustments on the
(Table 3). Fraction 2 presented the highest amount of all vo­ distillation system at industrial scale will be expected in
latiles analyzed (butanol, isoamyl alcohol, propanol, isobutyl order to obtain high quality cellulosic ethanol in accordance
alcohol, acetaldehyde, acetal, acetone, and ethyl acetate). with standard specifications. In this work, ethanol recovered
Such by-products mixtures (isoamyl alcohol (3-Methyl-1- by using batch equipment was important to analyze the
butanol), propanol, and isobutyl alcohol (2-Methyl-1-pro­ impact of the new substrate on the final product quality.
panol)) are called higher alcohols or fusel oils that can be
converted into valuable compound such as organic carbo­ 3.4. Vinasse characterization
nates (Bergmann et al., 2018). The excess of amino acids
during fermentation leads to amino acid deamination and Approximately 13.5 L of vinasse was generated as distillation
decarboxylation (Ehrlich Pathway) forming these higher al­ bottom product. Vinasse is an important side stream product
cohols (Walker and Stewart, 2016). Molasses is a source of that is currently used as fertilizer (Rossell et al., 2016), but it
amino acids (Alcantara et al., 2020); also, the cellulolytic en­ can be applied for the production of biogas (Fuess et al.,
zymes broth can contribute to the nitrogen content. Higher 2018). Table 5 presents the vinasse characterization in

Table 4 – Characterization of distilled fractions in comparison to standard specification of ANP


(National Agency of Petroleum - Brazilian Agency) needed to meet the fuel requirements sale.
Parameter Fraction 3 Fraction 4 ANP standard

Color < 20 < 20 5


Total acidic (mg/L) 29.3 51.1 30 (max.)
Electrical conductivity (µS/m) 312 980 300 (max.)
pH 5.6 4.9 6.0–8.0
Permanganate test (min) <1 <1 Not detected
Sodium content (mg/kg) 2.7 0.9 2.0 (max.)
Sulfur content (mg/kg) < 3.5 < 3.5 Not detected
Ethanol (% v/v) 79.5 43.9 92.5 – 94.6
Methanol (%v/v) < 0.03 < 0.03 0.5 (max.)
574 Chemical Engineering Research and Design 196 (2023) 568–576

Table 5 – Distillation bottom product (vinasse) characterization in comparison to vinasse obtained from ethanol
fermentation of cellulosic hydrolysate and sugarcane juice.
Compound/ parameter Value (this work) Value (Petrobras, 2012) Value (Petrobras, 2012)
Cellulosic hydrolysate Cellulosic hydrolysate Sugarcane juice

Calcium (mg/L) 404 ± 2 8–12 130–1540


Cupper (mg/L) 0.024 ± 0.001 - -
Iron (mg/L) 26 ± 1 - -
Magnesium (mg/L) 211 ± 1 16–24 200–490
Manganese (mg/L) 7.315 ± 0.004 - -
Molybdenum (mg/L) 0.230 ± 0.001 - -
Nickel (mg/L) 6.11 ± 0.08 - -
Potassium (mg/L) 1194 ± 1 40–80 1200–2100
Selenium (mg/L) < 0.005 - -
Sodium (mg/L) 10,752 ± 18 - -
Sulfate (mg/L) 24,870 ± 100 44–366 600–760
Zinc (mg/L) 0.500 ± 0.003 - -
Total nitrogen (mg NH3/L) 28.1 ± 0.4 - -
Total nitrogen (mg N/L) 108.60 ± 0.35 205 – 462 150–700
Total phenolic (mg/L) < 0.010 n.d. 0.4 – 12.4
Total phosphorous (mg/L) 128 ± 4 100.5 10–210
Total solids (mg/L) 62,490 ± 86 454 – 5805 23,700
Biochemical oxygen demand (mg/L) 25,049 ± 4 31,500–87,700 6000–16,500
Chemical oxygen demand (mg O2/L) 54,062 ± 26 75,800 – 109,700 15,000–33,000

Note: (-) means not determined and (n.d.) means not detected.

comparison to reported values (Petrobras, 2012) for the dis­ CRediT authorship contribution statement
tillation of fermented sugarcane juice and cellulosic hydro­
lysate. The vinasse originated from cellulosic ethanol can be Conceptualization, C.K.Y., R.M.F. and C.E.V.R.; methodology,
applied as fertilizer, because the content of phosphorous C.K.Y., S.T.R. and W.E.H.; investigation, C.K.Y.; data curation,
(128 ± 4 mg/L) and magnesium (211 ± 1 g/L) was similar to C.K.Y and S.T.R.; writing-reviewing and editing, C.K.Y.;
the current values. Also, it can be applied for biogas pro­ R.W.M. and R. M. F.; visualization, C.K.Y.; supervision, R.M.F.
duction, since the quantity of biochemical oxygen demand “All authors have read and agreed to the published version of
was high (25,049 ± 4 mg/L), also higher compared to the the manuscript.”
current values (6000 - 16,500 mg/L). However, sodium and
sulfate contents were higher than expected. Conflicts of interest

The authors declare no competing interests.


4. Conclusions

The fermentation of cellulosic hydrolysate obtained from


Acknowledgements
sugarcane bagasse by applying the current technologies was
This work used facilities at the National Laboratory of
well succeed with similar industrial performance. The char­
Biorenewables (LNBR), at the National Research Center for
acterization of the distilled products showed discrepancies
Energy and Materials (CNPEM), a Social Organization su­
compared to the standard values required for fuel, which can
pervised by the Ministry of Science, Technology, and
be solved by adjusting the operating parameter to remove all
Innovation (MCTI). We thank LNBR/CNPEM and UNICAMP for
impurities (volatile compounds). The bottom product pre­
providing the infrastructure and sharing expertise.
sented similar characteristics of the current vinasse from the
conventional process, except for sodium and sulfate con­
tents. Suitable equipment and operation of enzymatic hy­
References
drolysis can minimize the use of buffers; also other
Alcantara, G.U., Nogueira, L.C., Stringaci, L., de, A., Moya, S.M.,
substances can be used for pH control. Ideally, generating the
Costa, G.H.G., 2020. Brazilian “flex mills”: ethanol from su­
correspondent salt that will be metabolized by the micro­ garcane molasses and corn mash. Bioenergy Res. 13, 229–236.
organism (e.g., ammonium sulfate). However, for ethanol https://doi.org/10.1007/s12155-019-10052-3
production, yeast growth must be minimized. In conclusion, Alvarez, M.E.T., Bermúdez, J.H., De Moraes, E.B., Bonon, A.J., Wolf-
efforts should focus on improving enzymatic hydrolysis, Maciel, M.R., 2013. Heat transfer evaluation of multi­
such as reactors design and process strategies combined component batch distillation of the wine of sugarcane fer­
with more online measurements to maximize the solid-li­ mentation. Chem. Eng. Trans. 32, 517–522. https://doi.org/10.
3303/CET1332087
quid transference and accurate processing control.
Andrić, P., Meyer, A.S., Jensen, P.A., Dam-Johansen, K., 2010.
Reactor design for minimizing product inhibition during en­
zymatic lignocellulose hydrolysis. II. Quantification of inhibi­
Funding
tion and suitability of membrane reactors. Biotechnol. Adv. 28,
407–425. https://doi.org/10.1016/j.biotechadv.2010.02.005
This work was supported by the State of São Paulo Research Basso, L.C., Basso, T.O., Rocha, S.N., 2010. Ethanol production in
Foundation, FAPESP, Brazil, (grant numbers 2011/51902-9, Brazil: the industrial process and its impact on yeast fer­
2015/20630-4 and 2017/23335-9). mentation. Biofuel Prod. Recent Dev. Prospect. 1530, 85–100.
Chemical Engineering Research and Design 196 (2023) 568–576 575

Bergmann, J.C., Trichez, D., Sallet, L.P., de Paula e Silva, F.C., ethanol fermentations 1–9. https://doi.org/10.21203/rs.3.rs-
Almeida, J.R.M., 2018. Technological advancements in 1G 38002/v1.
ethanol production and recovery of by-products based on the Lischeske, J.J., Stickel, J.J., 2019. A two-phase substrate model for
biorefinery concept. adv. sugarcane biorefinery technol. enzymatic hydrolysis of lignocellulose: application to batch
commer. Policy issues paradig. Shift Bioethanol -Prod. 73–95. and continuous reactors. Biotechnol. Biofuels 12, 1–15. https://
https://doi.org/10.1016/B978-0-12-804534-3.00004-5 doi.org/10.1186/s13068-019-1633-2
Canilha, L., Chandel, A.K., Suzane dos Santos Milessi, T., Lopes, M.L., Paulillo, S.C., de, L., Godoy, A., Cherubin, R.A.,
Antunes, F.A.F., Luiz da Costa Freitas, W., das Graças Almeida Lorenzi, M.S., Giometti, F.H.C., Bernardino, C.D., de Amorim
Felipe, M., da Silva, S.S., 2012. Bioconversion of sugarcane Neto, H.B., de Amorim, H.V., 2016. Ethanol production in
biomass into ethanol: an overview about composition, pre­ Brazil: a bridge between science and industry. Braz. J.
treatment methods, detoxification of hydrolysates, enzymatic Microbiol 47, 64–76. https://doi.org/10.1016/j.bjm.2016.10.003
saccharification, and ethanol fermentation. J. Biomed. Mayer, F.D., Feris, L.A., Marcilio, N.R., Staudt, P.B., Hoffmann, R.,
Biotechnol. (2012), 1–15. https://doi.org/10.1155/2012/989572 Baldo, V., 2015. Influence of fusel oil components on the dis­
Charoenbhakdi, S., Dokpikul, T., Burphan, T., Techo, T., tillation of hydrous ethanol fuel (HEF) in a bench column.
Auesukaree, C., 2016. Vacuolar H+-ATPase protects Braz. J. Chem. Eng. 32, 585–593. https://doi.org/10.1590/0104-
Saccharomyces cerevisiae cells against ethanolinduced oxi­ 6632.20150322s00003215
dative and cell wall stresses. Appl. Environ. Microbiol. 82, Mizik, T., 2021. Economic aspects and sustainability of ethanol
3121–3130. https://doi.org/10.1128/AEM.00376-16 production—a systematic literature review. Energies. https://
Della-Bianca, B.E., Basso, T.O., Stambuk, B.U., Basso, L.C., doi.org/10.3390/en14196137
Gombert, A.K., 2013. What do we know about the yeast strains Narendranath, N.V., 2015. System for the treatment of biomass to
from the Brazilian fuel ethanol industry. Appl. Microbiol. facilitate the production of ethanol. US 9,034,620 B2.
Biotechnol. https://doi.org/10.1007/s00253-012-4631-x Olivieri, G., Wijffels, R.H., Marzocchella, A., Russo, M.E., 2021.
Dias, M.O.S., Modesto, M., Ensinas, A.V., Nebra, S.A., Filho, R.M., Bioreactor and bioprocess design issues in enzymatic hydro­
Rossell, C.E.V., 2011. Improving bioethanol production from lysis of lignocellulosic biomass. Catalysts 11, 1–19. https://doi.
sugarcane: evaluation of distillation, thermal integration and org/10.3390/catal11060680
cogeneration systems. Energy 36, 3691–3703. https://doi.org/ Palmqvist, E., Hahn-Hägerdal, B., 2000. Fermentation of lig­
10.1016/J.ENERGY.2010.09.024 nocellulosic hydrolysates. II: Inhibitors and mechanisms of
Dias, M.O.S., Junqueira, T.L., Cavalett, O., Cunha, M.P., Jesus, inhibition. Bioresour. Technol. 74, 25–33. https://doi.org/10.
C.D.F., Rossell, C.E.V., Maciel Filho, R., Bonomi, A., 2012. 1016/S0960-8524(99)00161-3
Integrated versus stand-alone second generation ethanol Petrobras, P.B.S.A., 2012. Method for producing energy-rich gases
production from sugarcane bagasse and trash. Bioresour. from lignocellulosic material streams. WO/2012/003556.
Technol. 103, 152–161. https://doi.org/10.1016/j.biortech.2011. Rivera, E.C., Yamakawa, C.K., Herrera, M., Geraldo, C., Rossell,
09.120 C.E.V., Maciel, R., Bonomi, A., 2013. A procedure for estimation
Dias, M.O.S., Junqueira, T.L., Rossell, C.E.V., MacIel Filho, R., of fermentation kinetic parameters in fed-batch bioethanol
Bonomi, A., 2013. Evaluation of process configurations for production process with cell recycle. Chem. Eng. Trans. 32,
second generation integrated with first generation bioethanol 1369–1374.
production from sugarcane. Fuel Process. Technol. 109, 84–89. Rivera, E.C., Yamakawa, C.K., Saad, M.B.W., Atala, D.I.P.,
https://doi.org/10.1016/j.fuproc.2012.09.041 Ambrosio, W.B., Bonomi, A., Junior, J., Rossell, C.E.V., 2017.
Ethanol Producer Magazine, 2021. Instant Intel [WWW Effect of temperature on sugarcane ethanol fermentation:
Document]. URL 〈https://ethanolproducer.com/articles/18560/ Kinetic modeling and validation under very-high-gravity fer­
instant-intel〉. mentation conditions. Biochem. Eng. J. 119, 42–51. https://doi.
Fuess, L.T., Garcia, M.L., Zaiat, M., 2018. Seasonal characterization org/10.1016/j.bej.2016.12.002
of sugarcane vinasse: assessing environmental impacts from Rossell, C.E.V., Yamakawa, C.K., Rivera, E.C., Nolasco Junior, J.,
fertirrigation and the bioenergy recovery potential through 2016. In: Peter Rein (Ed.), Ethanol production. Cane Sugar
biodigestion. Sci. Total Environ. 634, 29–40. https://doi.org/10. Engineering. Bartens, pp. 655–711.
1016/j.scitotenv.2018.03.326 Rossi, L.M., Gallo, J.M.R., Mattoso, L.H.C., Buckeridge, M.S.,
Guigou, M., Cabrera, M.N., Vique, M., Bariani, M., Guarino, J., Licence, P., Allen, D.T., 2021. Ethanol from sugarcane and the
Ferrari, M.D., Lareo, C., 2019. Combined pretreatments of eu­ brazilian biomass-based energy and chemicals sector. ACS
calyptus sawdust for ethanol production within a biorefinery Sustain. Chem. \ Eng. 9, 4293–4295. https://doi.org/10.1021/
approach. Biomass-.-. Convers. Biorefin. 9. https://doi.org/10. acssuschemeng.1c01678
1007/s13399-018-0353-3 Silva, V.F.N., Arruda, P.V., Felipe, M.G. a, Gonçalves, A.R., Rocha,
Hames, B., Ruiz, R., Scarlata, C., Sluiter, A., Sluiter, J., Templeton, G.J.M., 2011. Fermentation of cellulosic hydrolysates obtained
D., 2008. Preparation of samples for compositional analysis, by enzymatic saccharification of sugarcane bagasse pre­
National Renewable Energy Laboratory. treated by hydrothermal processing. J. Ind. Microbiol.
Jones, R.P., Greenfield, P.F., 1987. Ethanol and the fluidity of the Biotechnol. 38, 809–817. https://doi.org/10.1007/s10295-010-
yeast plasma membrane. Yeast 3, 223–232. https://doi.org/10. 0815-5
1002/yea.320030403 Sluiter, A., Ruiz, R., Scarlata, C., Sluiter, J., Templeton, D., 2008.
Klein, B.C., Silva, J.F.L., Junqueira, T.L., Arruda, P.V., Ienczak, J.L., Determination of extractives in biomass. Natl. Renew. Energy
Mantelatto, P.E., Pradella, J.G., da, C., Vaz Junior, S., Bonomi, Lab https://doi.org/NREL/TP-510-42621.
A., 2017. Process development and techno- economic analysis Sluiter, A., Hames, B., Ruiz, R., Scarlata, C., Sluiter, J., Templeton,
of bio-based succinic acid derived from pentoses integrated to D., Crocker, D., 2012. Determination of structural carbohy­
a sugarcane biorefinery. Biofuels. Bioprod. Bioref. 6, 246–256. drates and lignin in biomass. Natl. Renew. Energy Lab https://
https://doi.org/10.1002/bbb.1813 doi.org/NREL/TP-510-42618.
Lee, S.S., Robinson, F.M., Wang, H.Y., 1981. Rapid determination Soccol, C.R., Vandenberghe, L.P.D.S., Medeiros, A.B.P., Karp, S.G.,
of yeast viability. Biotechnol. Bioeng. Symp. 11. Buckeridge, M., Ramos, L.P., Pitarelo, A.P., Ferreira-Leitão, V.,
Letti, L.A.J., Sydney, E.B., de Carvalho, J.C., de Souza Gottschalk, L.M.F., Ferrara, M.A., Silva Bon, E.P., Da, Moraes,
Vandenberghe, L.P., Karp, S.G., Woiciechowski, A.L., Soccol, V. L.M.P., De, Araújo, J.D.A., Torres, F.A.G., 2010. Bioethanol from
T., Novak, A.C., Magalhães Junior, A.I., Martinez Burgos, W.J., lignocelluloses: status and perspectives in Brazil. Bioresour.
de Carvalho Neto, D.P., Soccol, C.R., 2022. Roles and impacts of Technol. 101, 4820–4825. https://doi.org/10.1016/j.biortech.
bioethanol and biodiesel on climate change mitigation. Bi. 2009.11.067
Lino, F., Bajic, D., Vila, J., Sanchez, A., Sommer, M., 2020. Complex de Souza Dias, M.O., Maciel, R., Mantellato, P.E., Cavalett, O.,
yeast-bacteria interactions affect the yield of industrial Rossell, C.E.V., Bonomi, A., Leal, M.R.L.V., 2015. Sugarcane
576 Chemical Engineering Research and Design 196 (2023) 568–576

processing for ethanol and sugar in Brazil. Environ. Dev 15, fermentation of cellulosic hydrolysate from sugarcane ba­
35–51. gasse to develop stand-alone second generation ethanol
Townsend, T.J., Sparkes, D.L., Wilson, P., 2017. Food and bioe­ plant. Chem. Eng. Trans. 50, 163–168. https://doi.org/10.3303/
nergy: reviewing the potential of dual-purpose wheat crops. CET1650028
GCB Bioenergy 9, 525–540. https://doi.org/10.1111/gcbb.12302 Yamakawa, C.K., Atala, D.I.P., Ambrosio, W.B., Nolasco Junior, J.,
Volpi, M.P.C., Magalh, L.O., Winck, V., Labate, M.T.V., Moraes, B.S., Rossell, C.E.V., 2017. Assessment of VHG ( Very High Gravity)
2022. Metaproteomic Analysis of the Anaerobic Community ethanol fermentation in continuous multistage with cell re­
Involved in the Co-Digestion of Residues from Sugarcane covery, reactivation and recycling using a blend of juice and
Ethanol Production for Biogas Generation. molasses from sugarcane as raw material. Sugar Ind. 4,
Walker, G.M., Basso, T.O., 2020. Mitigating stress in industrial 212–213.
yeasts. Fungal Biol. 124, 387–397. https://doi.org/10.1016/j. Yamakawa, C.K., Ccopa Rivera, E., Kwon, H., Herrera Agudelo,
funbio.2019.10.010 W.E., Saad, M.B.W., Leal, J., Rossell, C.E.V., Bonomi, A., Maciel
Walker, G.M., Stewart, G.G., 2016. Saccharomyces cerevisiae in Filho, R., 2019. Study of influence of yeast cells treatment on
the production of fermented beverages. Beverages 2, 1–12. sugarcane ethanol fermentation: operating conditions and
https://doi.org/10.3390/beverages2040030 kinetics. Biochem. Eng. J. 147, 1–10. https://doi.org/10.1016/j.
Yamakawa, C.K., 2016. Avaliação da fermentação alcoólica com bej.2019.03.022
reciclo de células de hidrolisado celulósico de bagaço de cana- Zabed, H., Sahu, J.N., Suely, A., Boyce, A.N., Faruq, G., 2017.
de-açúcar em unidade integrada e autônoma. UNICAMP. Bioethanol production from renewable sources: current per­
Yamakawa, C.K., Rojas, S.T., Geraldo, V.C., Rivera, E.C., Herrera, spectives and technological progress. Renew. Sustain. Energy
W.E., Bonomi, A., Rossell, C.E.V., Maciel Filho, R., 2016. Ethanol Rev. https://doi.org/10.1016/j.rser.2016.12.076

You might also like