Bioethanol Author
Bioethanol Author
Bioethanol Author
Author Copy
Author Copy
Edited by
Ayerim Y. Hernández Almanza, PhD
Nagamani Balagurusamy, PhD
Héctor Ruiz Leza, PhD
Cristóbal N. Aguilar, PhD
Author Copy
For permission to photocopy or use material electronically from this work, access www.copyright.com or contact the Copyright
Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. For works that are not available on CCC
please contact mpkbookspermissions@tandf.co.uk
Trademark notice: Product or corporate names may be trademarks or registered trademarks and are used only for identification
and explanation without intent to infringe.
Author Copy
Prof. Hernández Almanza’s experience is in extraction and purification of
bioactive compounds, fermentation process, microbial pigments production,
and biological characterization. She is recognized as a member (candidate)
of the National System of Research (SNI) by the National Council of Science
and Technology of the Government of México. In 2019 she obtained the
“Mujer Universitaria” award in Scientific Ambit given by the Autonomous
University of Coahuila. Prof. Hernández-Almanza has published over 10
original research papers in indexed journals and 12 book chapters and has
participated and contributed in more than five scientific meetings. She has
realized two research stays: at Università degli Studi di Perugia, Perugia,
Italia (2013), and at Gachon University, Seongnam, Gyeonggi-do, South
Korea (2016).
Author Copy
the biomass pretreatment stage in the Cluster of Bioalcoholes in the Mexican
Centre for Innovation in Bioenergy (Cemie-Bio) from the Secretary of
Energy in Mexico. Dr. Ruiz obtained his PhD in Chemical and Biological
Engineering from the Center of Biological Engineering at the University of
Minho, Portugal, in 2011 and was a postdoctoral researcher at the University
of Minho (Portugal) and University of Vigo (Spain).
Dr. Ruiz is currently working on the development of biorefinery strategies
for the production of high added-value compounds and biofuels from ligno-
cellulosic, micro, and macroalgal biomass in the context of the bioeconomy.
He is technically responsible for several research projects (CONACYT) and
a technical consultant for biomass conversion companies in Mexico and
Europe. Dr. Ruiz has conducted several research stays and technical visits at
the University of North Texas (USA), Federal University of Sergipe (Brazil),
Brazilian Bioethanol Science and Technology Laboratory-CTBE (Brazil),
Chemical and Biological Engineering Department at the University of British
Columbia (Canada), CIEMAT-Renewable Energy Division, Biofuels Unit
(Spain), University of Jaén (Spain), Sadar Swaran Singh National Institute
of Bio-Energy (India), Tokyo Institute of Technology (Japan), University
of Concepción (Chile), Umeå University (Sweden), National Laboratory
of Energy and Geology-LNEG (Portugal), CSIR-National Institute for
Interdisciplinary Science and Technology (India), University of Florida-Stan
Mayfield Biorefinery Plant (USA), University of Kannur (India), Federal
University of Rio Grande do Norte (Brazil).
He has authored or co-authored 75 publications, including papers in
indexed journals and chapters, with over 2300 citations and an h-index of
28 (Google Scholar Citations). Currently, Dr. Ruiz is Editor-in-Chief of
BioEnergy Research Journal (Springer Publishing) and Associate Editor of
Biotechnology for Biofuels (BioMed Central-Part of Springer Nature), and
Author Copy
Level II distinction).
Contributors..............................................................................................................xi
Abbreviations......................................................................................................... xvii
Acknowledgment..................................................................................................... xxi
Preface..................................................................................................................xxiii
Author Copy
1. Physiology of Ethanol Production by Yeasts.................................................1
Miriam Soledad Valenzuela Gloria, Diana Laura Alva-Sánchez,
M. P. Luévanos Escareño, Cristóbal N. Aguilar, Nagamani Balagurusamy,
and Ayerim Hernández-Almanza
Author Copy
Iosvany López-Sandin, Francisco Zavala-García, Guadalupe Gutiérrez-Soto,
and Héctor Ruiz Leza
Index......................................................................................................................501
P. Abdeshahian
Biopolymers, Bioprocesses, Process Simulation Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
Cristóbal N. Aguilar
Bioprocesses and Bioproducts Research Group, BBG-DIA, Food Research Department,
School of Chemistry, Autonomous University of Coahuila (UAdeC), Saltillo–25280, Coahuila, Mexico
Author Copy
Diana Laura Alva-Sánchez
School of Biological Science, Autonomous University of Coahuila, Torreón–27000, Coahuila, Mexico
Lorena Amaya-Delgado
Industrial Biotechnology Unit, Center for Research and Assistance in Technology and Design of the State
of Jalisco A.C., Camino Arenero–1227, El Bajio del Arenal, C.P.–45019, Zapopan, Jalisco, Mexico
Melchor Arellano-Plaza
Industrial Biotechnology Unit, Center for Research and Assistance in Technology and Design of the State
of Jalisco A.C., Camino Arenero–1227, El Bajio del Arenal, C.P.–45019, Zapopan, Jalisco, Mexico
L. G. De Arruda
Biopolymers, Bioprocesses, Process Simulation Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
D. B. Arya
Department of Environmental Sciences, University of Kerala, Thiruvananthapuram, Kerala, India
Nagamani Balagurusamy
Bioremediation Laboratory, Biological Science Faculty, School of Biological Science,
Autonomous University of Coahuila (UAdeC), Torreón–27000, Coahuila, Mexico
T. R. Balbino
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
F. G. Barbosa
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
M. M. Campos
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
A. S. Cardoso
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
L. T. Carvalho
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology, Engineering School
of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
Tecnologico de Monterrey, Escuela de Ingenieria y Ciencias, Av. General Ramon Corona No. 2514,
Zapopan–45201, Jal., Mexico
M. J. Castro-Alonso
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
Author Copy
M. L. Silva da Cunha
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), 12.602.810, Lorena, SP, Brazil
Ileana Mayela María Moreno Dávila
Department of Biotechnology, School of Chemistry, Autonomous University of Coahuila, Mexico
Ahmet Demir
Department of Environmental Engineering, Yildiz Technical University, Turkey, Davutpasa Campus,
Istanbul–34220, Turkey
Kakoli Dutt
Department of Bioscience and Biotechnology, Banasthali Vidyapith, Rajasthan–304022, India,
Tel.: 9887398384, E-mail: kakoli_dutt@rediffmail.com
M. P. Luévanos Escareño
School of Biological Science, Autonomous University of Coahuila, Torreón–27000, Coahuila, Mexico
Norma M. De La Fuente-Salcido
Faculty of Biological Sciences, Autonomous University of Coahuila, Torreón, Coahuila, México,
E-mail: normapbr322@gmail.com
Anne Gschaedler
Industrial Biotechnology Unit, Center for Research and Assistance in Technology and Design of the State
of Jalisco A.C., Camino Arenero–1227, El Bajio del Arenal, C.P.–45019, Zapopan, Jalisco, Mexico
Guadalupe Gutiérrez-Soto
University of Nuevo León, Faculty of Agronomy, Francisco Villa S/N Col. Ex Hacienda
El Canadá–66415, General Escobedo, N. L., México
Ayerim Hernández-Almanza
Apple Academic Press
Héctor Hernández-Escoto
Chemical Engineering Department, University of Guanajuato, Noria Alta s/n, Guanajuato, 36050, Mexico
R. T. Hilares
Author Copy
Material Laboratory, Catholic University of Santa Maria (UCSM), Yanahuara, AR–04013, Peru
Subburamu Karthikeyan
Department of Renewable Energy Engineering, Tamil Nadu Agricultural University, Coimbatore,
Tamil Nadu, India, E-mail: skarthy@tnau.ac.in
David Francisco Lafuente-Rincón
Bioprocesses and Bioprospecting Laboratory, Biological Sciences Faculty,
Autonomous University of Coahuila, Torreón–27000, Coahuila, México
Juan A. León
Laboratory of Process Intensification and Hybrid System (LIPSH), Department of Chemical Engineering,
National University of Colombia, Manizales, Caldas, Colombia, E-mail: jaleonma@unal.edu.co
Iosvany López-Sandin
University of Nuevo León, Faculty of Agronomy, Francisco Villa S/N Col. Ex Hacienda
El Canadá–66415, General Escobedo, N. L., México
P. R. F. Marcelino
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
E. Mier-Alba
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
Bestami Ozkaya
Department of Environmental Engineering, Yildiz Technical University, Turkey, Davutpasa Campus,
Istanbul–34220, Turkey
Alan D. Perez
Apple Academic Press
Sustainable Process Technology (SPT), University of Twente, Enschede, Overijssel, The Netherland,
E-mail: a.d.perezavila@utwente.nl
Laura Andrea Pérez-García
Faculty of Biological Sciences, Autonomous University of Coahuila, Torreón, Coahuila, México
César D. Pinales-Márquez
Biorefinery Group, Food Research Department, Faculty of Chemistry Sciences,
Autonomous University of Coahuila, Saltillo–25280, Coahuila, Mexico
C. A. Prado
Author Copy
Biopolymers, Bioprocesses, Process Simulation Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
Javier Quintero
Department of Chemical Engineering, National University of Colombia, Manizales, Caldas, Colombia,
E-mail: jaquinteroj@unal.edu.co
Rosa M. Rodríguez-Jasso
Biorefinery Group, Food Research Department, Faculty of Chemistry Sciences, Autonomous University
of Coahuila, Saltillo–25280, Coahuila, Mexico, Phone: (+52)-1-844-416-12-38,
E-mail: rrodriguezjasso@uadec.edu.mx
Teresa Romero-Gutiérrez
Computer Sciences Department, Exact Sciences and Engineering University Centre, Universidad de
Guadalajara, Blvd. Gral. Marcelino Garcia Barragan–1421, Olimpica. Guadalajara, Mexico
D. Rubio-Ribeaux
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil
S. Sánchez-Muñoz
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil,
E-mails: salvador.sanchez@usp.br; sanchezmunoz.ssm@gmail.com
Dania Sandoval-Nuñez
Industrial Biotechnology Unit, Center for Research and Assistance in Technology and Design of the State
of Jalisco A.C., Camino Arenero–1227, El Bajio del Arenal, C.P.–45019, Zapopan, Jalisco, Mexico
J. C. Santos
Biopolymers, Bioreactors, and Process Simulation Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), 12.602.810. Lorena, SP, Brazil,
E-mail: jsant200@usp.br
Apple Academic Press
Charu Saraf
Department of Bioscience and Biotechnology, Banasthali Vidyapith, Rajasthan–304022, India,
Tel.: 9910331924, E-mail: charusaraf17@gmail.com
Rohit Saxena
Biorefinery Group, Food Research Department, Faculty of Chemistry Sciences,
Autonomous University of Coahuila, Saltillo–25280, Coahuila, Mexico
Shiva
Biorefinery Group, Food Research Department, Faculty of Chemistry Sciences,
Author Copy
Autonomous University of Coahuila, Saltillo–25280, Coahuila, Mexico
S. S. Da Silva
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology,
Engineering School of Lorena, University of São Paulo (EEL-USP), 12.602.810. Lorena, SP, Brazil,
E-mail: silviosilverio@gmail.com
R. T. Terán-Hilares
Material Laboratory, Catolic University of Santa Maria (UCSM), Yanahuara–04013, AR, Perú
Oznur Yildirim
Department of Environmental Engineering, Yildiz Technical University, Turkey, Davutpasa Campus,
Istanbul–34220, Turkey
Francisco Zavala-García
University of Nuevo León, Faculty of Agronomy, Francisco Villa S/N Col. Ex Hacienda
El Canadá–66415, General Escobedo, N. L., México, E-mail: francisco.zavala.garcia@gmail.com
Author Copy
AFEX ammonia fiber expansion
ALDH acetaldehyde by aldehyde dehydrogenase
ARP ammonia recycling percolation
ATPS aqueous two-phase systems
BFP biofuture platform
C4H cinnamate 4-hydroxylase
CBP consolidated bioprocessing
CCCF closed-circulating fermentation system
CEMIE-Bio Mexican Center for Innovation in Bioenergy
CEs carbohydrate esterases
CMFS continuous membrane fermenter separator
CNOOC China National Offshore Oil Corporation
CO2 carbon dioxide
COMT caffeic acid 3-O-methyltransferase
CRISPR clustered regularly interspaced short palindromic repeats
CS chitosan
CSIR Industrial Scientific Research Council
CSTR continuous stirred tank reactors
CWI cell wall integrity
DDGs distillers dried grains
DDG dry distillation grain
DESs deep eutectic solvents
DHAP dihydroxyacetone phosphate
DMC direct microbial conversion
DP degree of polymerization
DSSF delayed simultaneous saccharification and fermentation
DyP dye decolorizing peroxidase
EA ethanol adaption
ED Entner-Doudoroff
ED extractive distillation
EISA energy independence and security act
EMP Embden-Meyerhof-parnas
EPA energy policy act
Apple Academic Press
EPS exopolysaccharides
ESR environmental stress response
EtOH alcohol
EU European Union
FISH fluorescent in situ hybridization
FST fiber separation technology
G3P glycerol-3-phosphate
Author Copy
GAP glyceraldehyde-3-phosphate
GBEP global bioenergy partnership
GHG greenhouse gas
GRAS generally recognized as safe
gTME global transcription machinery engineering
GW global warming
H2O2 hydrogen peroxide
HAA 3-hydroxyanthranilic acid
HAD homogenous azeotropic distillation
HC hydrodynamic cavitation
HCl hydrochloric acid
HCR hydrodynamic cavitation reactors
HCT hydroxycinnamoyl transferase
HER hydrogen evolution reaction
HMF 5-hydroxymethylfurfural
HOG high osmolarity glycerol
HPLC high-performance liquid chromatography
HR homologous recombination
HSP heat shock proteins
HTAD heterogeneous azeotropic distillation
HTLR hydrothermal liquefaction reactors
HyPol hyperbranched polymer
iHG integrated high gravity
IL ionic liquids
IMHeRE intelligent microbial heat-regulating engine
ISC iron-sulfur clusters
ISPR in situ product recovery
KDPG 2-keto-3-deoxy-6-phosphogluconic acid
Author Copy
MEDC microbial electrodialysis cell
MES microbial bioeletroshyntesis
MFC microbial fuel cell
MnP manganese peroxidase
MREC microbial reverse electrodialysis electrolysis cell
MWIR microwave irradiation reactor
MWR microwave reactors
NGS next-generation sequencing
NRC national research council
OD oven-dried
OPEC Organization of Arab Petroleum Exporting Countries
PA phosphatidic acid
PDC pyruvate decarboxylase
PDMS polydimethylsiloxane
PFD prefoldin
Pgd phosphogluconate dehydrogenase
PKA protein kinase A
PM particle matters
PPP pentose phosphate pathway
PSA pressure swing adsorption
PSSF pre-saccharification and fermentation strategy
PTA phosphotransacetylase
PTMSP poly(1-trimethylsily-1-propyne)
PV pervaporation
RDB renewable diesel blendstock
REFINE RNA seq examiner for phenotype-informed network engineering
RFS renewable fuel program
RGI rhamnogalacturonan I
Author Copy
SMT selective milling technology
sRNA small RNA
SSCF simultaneous saccharification and co-fermentation
SSF simultaneous saccharification and fermentation
STBR stirred tank bioreactors
STR stirred tank reactor
Tal transaldolase
TALEN’s transcription activator-like effector nuclease
TAME TALEN assisted multiplex editing
TEM transmission electron microscopy
TEMPOL 4-Hydroxy-2,2,6,6-tetramethylpiperidin-1-oxyl
TF transcription factors
TiO2 titanium dioxide
TKL1 transketolase
TPP trehalose-6-phosphate phosphatase
TPS trehalose-6-phosphate synthase
VP versatile peroxidase
WW water washing
XDH xylitol dehydrogenase
XGA xylogalacturonan
XI xylose isomerase
XOs xylooligosaccharides
XR xylose reductase
XYL1 xylose reductase heterologous
ZFN zinc finger nuclease
Author Copy
• Raúl Rodríguez Herrera
Autonomous University of Coahuila, Mexico
• Arturo Sócrates Palacios
ESPOL Polytechnic University, Ecuador
• Leopoldo Javier Ríos González
Autonomous University of Coahuila, Mexico
• Alejandra Alvarado
University of Tübingen, Germany
• Danay Carrillo Nieves
Tecnológico de Monterrey, Mexico
• Aldo Ricardo Almeida Robles
University of Copenhagen, Denmark
• Miguel Ángel Medina Morales
Autonomous University of Coahuila, Mexico
Author Copy
biological potential. A clear example is the use of corn and sugarcane
bagasse, rich in sugars, to obtain bioethanol. Bioethanol is one of the most
interesting biofuels since it has a positive characteristic on the environment.
During the last decades, there has been a great interest in the production
and use of biodiesel or bioethanol as promising alternatives to replace fossil
fuels. For this reason, the development of biorefineries and processes, as well
as the design of bioreactors, have been fundamental issues to understand this
area. Likewise, the study of metabolic and physiological processes that are
carried out by some bioethanol-producing microorganisms is also necessary
to know. Yeast, like Zymomonas mobilis and Clostridium thermocellum,
have been demonstrated to have a prominent role in bioethanol production.
However, it is necessary to delve into factors such as increasing the tolerance
of these strains to produce bioethanol, improve genetic regulation and
implement genetic engineering methods that favor production yields.
On the other hand, the development of tools and strategies that allow
obtaining bioethanol of major quality, viable alternatives for purification of
the product are factors that still have to be improved. Bioethanol production
has proven to be a promising option to reduce the damage caused by the
conventional processes used to obtain fuels; therefore, there are still some
challenges to overcome.
This book includes the advances and perspectives within the bioethanol
industry, also describes some biochemical and physiological parameters
carried out by the main bioethanol producing microorganisms as well as the
potential applications that this bioproduct can have and the advantage that it
would generate.
—Editors
Author Copy
CRISTÓBAL N. AGUILAR,2 NAGAMANI BALAGURUSAMY,1 and
AYERIM HERNÁNDEZ-ALMANZA1
1
School of Biological Science, Autonomous University of Coahuila,
Torreón–27000, Coahuila, Mexico,
E-mail: ayerim_hernandez@uadec.edu.mx (A. Hernández-Almanza)
2
Bioprocesses and Bioproducts Research Group, BBG-DIA, Food Research
Department, School of Chemistry, Autonomous University of Coahuila,
Saltillo–25280, Coahuila, Mexico
ABSTRACT
Yeasts have been gaining popularity due to their ability to produce a series
of compounds of interest such as pigments, phenolic compounds, fatty acids,
enzymes, and even under the right conditions, and they are ethanol producers.
The latter is characterized by producing ethanol and carbon dioxide (CO2)
under anaerobic conditions using fermentable sugars as substrates. To
consider that yeast is suitable for the production of ethanol, it is desirable
that it meets some characteristics of adaptability to the use of different
sources of carbon and nitrogen, to acidity or low availability of glycerol, and
even tolerance to high levels of ethanol concentration. In addition to these
criteria, it is also important to consider that the fermenting strain must be
able to use hexose and pentose and tolerate the inhibitory by-products of the
pretreatment. In such a way that by manipulating both carbon and nitrogen
sources, as well as environmental factors, it would be possible to increase
or decrease ethanol production, which requires knowledge of the species to
be used. This is widely used in the industry, in which over time, the use of
ethanol has been diversifying more and more, so what began to be used in
the food industry for the production of alcoholic beverages, today it is used
for the production of sustainable alternative fuels, which have gained impact
in recent years thanks to their profitable production thanks to the low cost of
Apple Academic Press
1.1 INTRODUCTION
Author Copy
in various industrial fermentation processes due to their ability to convert high
concentrations of sugars into ethanol and CO2; for example, Saccharomyces
cerevisiae has been exploited for centuries for the production of alcoholic
beverages. Currently, ethanol obtained from this via is an alternative to be
used as a substitute for gasoline [1–3].
The ethanol that is produced through fermentation represents a positive
alternative as fuel to petroleum, as a source of energy for batteries through
electrochemical effects, as energy in the generation of energy by thermal
combustion, etc. [4]. Ethanol is significantly less toxic to humans than is
gasoline in such a way that it reduces air pollution thanks to its low volatility,
photochemical, and waste activity [5]. Ethanol obtained from waste materials
from biomass or renewable sources is called bioethanol, and it can be used as
fuel, chemical raw material, and solvent in various industries.
The production of ethanol as liquid fuel is obtained by fermentation
from biomass, sugar cane, cereal grains, and sugar beets, and it is gaining
a lot of popularity around the world [4, 6]. If you have greater control of
the fermentation conditions, this can contribute to the reduction of stress
towards yeast cells and decrease contamination by bacteria and wild yeasts.
Therefore, having large information gaps around the ethanol production
processes, leading to large investigations in order to achieve the generation
of higher quality products and more optimized processes [7].
Author Copy
as substrate rich in sugars for the production of ethanol (Table 1.1). For
example, yeasts such as Schizosaccharomyces pombe, Candida krusei,
Kluyveromyces marxianus, Dekkera bruxellensis, Pichia striptis, Pichia
kudriavzevii, Wickerhamomyces anomalous, among others; they have been
isolated and identified as producing ethanol with good behavior and tolerant
to high concentrations of alcohol (EtOH) [17–19].
The main factors that contribute to the low growth of yeasts in the process
and that lead to high ethanol yields characterized in 90–92% of the theoretical
conversion of sugar to ethanol are high cell densities, cell recycling, and high
ethanol concentration [20–22].
Author Copy
Physiology of Ethanol Production by Yeasts 5
Author Copy
FIGURE 1.1 Rough diagram of the production of ethanol from lignocellulosic materials.
Author Copy
respiration. In this process, a direct competition is generated in the action of
PDC, between its use for the production of ethanol and the maintenance of
the balance of pyruvate availability in the metabolic pathway. Subsequently,
acetaldehyde is converted to ethanol by an alcohol dehydrogenase (ADH)
[13, 25]. Due to its nature as an oxidoreductase it manages to promote the
reversible interconversion of alcohols and the aldehydes/ketones themselves
[25]. ADH has available a number of substrates present in the different
metabolic pathways, which results in the need for a strict order in order to
preserve the homeostasis of the intermediates and products [1]. Therefore,
this is why eukaryotes and even humans have numerous ADH enzymes. Of
which it is possible to mention the enzymes Adh1, Adh2 and Adh3 [11].
The Adh1 enzyme is crucial during fermentation, since it replenishes the
NAD + assembly in order to produce ethanol; In addition, glucose manages
to suppress the enzyme Adh2, thus allowing the oxidation of ethanol, this
happens only under certain conditions of need. And last but not least, Adh3 is
expressed in the mitochondria allowing to carry out its main function which
is to maintain the redox balance of the process [1, 25, 29].
sugars as glucose or fructose can be easily harnessed by yeasts since these are
assimilated right after being hydrolyzed either inside or outside the cell [33].
For the growth process of yeasts not all sugars are used simultaneously,
the process is started with the fermentation of sucrose with the help of the
enzyme invertase, they cause a hydrolysis allowing the generation of glucose
and fructose for the use of yeast, during the initiation phase within fermenta-
tion, glucose is used in a greater proportion than fructose [34], due to the
Author Copy
glycolytic property of yeast. When the values of these saccharides decrease
under among concentration threshold maltose consumption increases if its
available, which is variable due to its entry into the cell thanks to the action
of maltose-permease enzyme; when it is present, it proceeds to the gradual
use of maltotriose [34, 35].
To mention some of the yeasts capable of fermenting maltotriose, maltose,
sucrose, glucose, and even fructose, they are Saccharomyces cerevisiae and
Saccharomyces uvarum. The importance of fermentable sugars derives from
their usefulness for the production of ethanol, CO2, and some representative
amino acids [36, 37].
The presence of carbon sources in a critical level turns cells metabolism
from respiratory growth into fermentative growth. Among of fermentable
sugars, glucose is the one who suppresses the enzyme that intervenes in the
metabolism of other sugars and in respiratory growth. Then when glucose
finds at critic concentration is converted into ethanol and CO2, this in spite of
a surplus of air; the optimal range for this to occur would be from 35 to 280
mg/L, which is equivalent to 5% of concentration [35, 38, 39].
Over the course of fermentation, a small amount of oxygen is present
in the broth, which goes hand in hand with the tolerance to ethanol and the
viability of the cell, a fact that allows the synthesis of sterols and is related to
fatty acids unsaturated around the membrane [30, 35].
In a broader perspective on the carbon sources, there may be decomposed
into four different sections; starting with the simple disaccharides also
known as hexoses, continuing with the starches, mainly of vegetable origin,
to this triad are also joined the lignocellulosic materials, and finally the agro-
industrial waste [25, 33, 36]. It is important to mention that each one has its
own pros and cons, for example, in terms of simple sugar-based materials,
they can be easily and immediately used as the sole substrate by microbial
strains capable of producing ethanol [25]. On the other hand, due to the excess
of agricultural production, the use of starch for the production of ethanol has
emerged, however, it is recognized that the levels of performance of this are
too low to achieve the absolute replacement of gasoline [36]. After this is
mandatory to mention the use of lignocellulosic materials as very rentable
Apple Academic Press
source for its low cost and easy obtention, without leaving aside the energy
required to obtain the necessary sugars from LCB, it is important to mention
that the main disadvantage is the availability of said substrates [25, 36].
Furthermore, the complexity of the substrate is what will dictate whether
a hydrolysis is used, either enzymatic or acidic, or if it is worked in a crude
way, the latter being applicable to agro-industrial waste. Therefore, for this
to suck, it is necessary to provide a rich source of nitrogen, since not all
Author Copy
yeasts are capable of using mineral nitrogen [33, 34]. Even within these there
is a classification; first, we have the sources of mineral nitrogen where both
ammonium salts and urea are located, as they have an assimilation capacity
almost on par; second, there are the sources of organic nitrogen, where the
presence of glutamic and aspartic acids are noted, as well as their amines
(asparagine) in both D and L forms [35]. The aim of nitrogen present in
nitrogenous compounds is that it dissociates in the form of ammonium
ions and hydrogens that provide alkaline pH values to the growth medium,
achieving this by adding such compounds.
It is important to mention that some yeasts and even filamentous fungi
use nitrate as a source of nitrogen, whose assimilation happens thanks
to the reduction to ammonia due to the action of the enzyme ammonium
reductase [40].
Thus, at the time when the yeasts are subjected to certain light doses of
tension, some of their cells express a characteristic called cross-protection,
with which they become resistant to large and generally lethal doses of
other tensions. This characteristic is developed due to the activation of a
Apple Academic Press
Author Copy
and vitality, which are defined as the ability to reproduce and the metabolic
activity of a culture, respectively. So, for these two conditions to occur during
propagation, the participation of storage carbohydrates, such as glycogen and
trehalose, is necessary. Glycogen serves as the main carbohydrate that stores
energy in yeast allows the synthesis of sterols, trehalose, and fatty acids,
throughout the delay stage. Due to this, it is desirable that the growth media
contain high concentrations of said polysaccharide so that the synthesis of
sterols and fatty acids is carried out, which represent a high expenditure of
energy. Glycogen accumulation is the consequence of the limited availability
of nitrogen or carbon. As for trehalose, it is considered to be a protective
oxidative disaccharide in situations of cellular stress such as high levels of
osmolarity, decreased nutrients, hunger, high, and low temperatures, and high
concentration of ethanol, which in high levels intracellular concentration
allows cell viability through the initial stages of fermentation and, therefore,
increases the rates of carbohydrate utilization [42, 44].
Temperature is the main factor that affects the production of ethanol, due
to the fermentation process is monitored based on the development of heat,
that is, the increase in heat energy in response to the metabolic activities
of microorganisms. One of the ways of inducing heat stress would be the
absence of cooling, which would lead to reduced growth of microorganisms
and fermentation defects. Therefore, the production of ethanol in tropical
countries has been managed at high temperatures in order to achieve a
reduction in costs in the cooling process. In addition, high temperature
management during fermentation provides a number of advantages, such
as efficient saccharification and fermentation, a constant change from
Author Copy
that in general most species grow very well around 25°C [30].
As mentioned in paragraphs preceding high temperature stress (or heat
shock) in yeast cells, heat damage can alter hydrogen bonds and even
hydrophobic interactions, leading to the unfolding of proteins and nucleic
acids. Therefore, anaerobic fermentation, being an exothermic reaction,
ends up increasing the temperature in some species, which are known
as thermotolerant, thus reaching fermentation temperatures above 40°C.
The term “thermotolerant” is used for those yeasts that have the transient
ability in their cells to survive subsequent lethal exposures at elevated
temperatures, that is, after a sudden thermal shock. Heat shock responses
in yeast occur when cells move rapidly at elevated temperatures, and if
this is near-fatal, it will produce “heat shock proteins (HSPs)” with a high
level of conservation, in response to the induced synthesis of a specific
set of proteins. HsP perform numerous physiological functions, including
thermo-protection [11, 30, 41, 44].
1.5.2 INFLUENCE OF pH
Author Copy
the aeration of the must, since the presence of oxygen in the initial fermen-
tation phase is essential for the rapid reproduction of yeast and the complete
restoration of sugar. By maintaining the correct amount of glycogen and
trehalose, the correct aeration improves the vitality of the biomass and the
immunity of the yeast cells to different tensions; carrying out in this way a
selective exchange of metabolites and a correct extraction of nutrients from
the environment. When a low amount of oxygen is present in the medium,
a delay in fermentation or a change in the sensory properties of the product
is caused [48].
The aeration of a culture has the advantage of providing the necessary
oxygen for respiration and the elimination of CO2 produced by the metabolism
of carbonate compounds. Therefore, the supply of oxygen through aeration
must be sufficient to avoid the generation of alcohol (EtOH) throughout
the fermentation process and it is recommended that the added air be rich
in oxygen. Then, during growth and metabolism, the transfer of nutrients,
and solid, solid, and gaseous metabolites between the environment and the
cell is essential and continuous, this is where the solubility of the gas enters
the liquid phase, so that it allows the exchange between phases. liquid and
gaseous [35, 49, 52].
Yeasts require appropriate amounts of oxygen to carry out their oxidative
metabolism and thus proceed with the optimal production of ethanol; by
using this gas, they are able to synthesize unsaturated fatty acids and sterols,
which are necessary for continuous anaerobic growth and cell division.
Even S. cerevisiae, which is anaerobic in growth, requires small amounts of
molecular oxygen for the synthesis of fatty acids and sterols. Furthermore,
pentose fermentation yeasts require very small amounts of oxygen, which
must be constantly monitored. So, since the fermentation rate is proportional
to the number of metabolically active yeasts present, this means that the
attenuation of the medium goes hand in hand with the availability of oxygen
Author Copy
process, but also improves the final characteristics of the product. Yeast with
low oxygen requirements generally produces a slightly higher number of
esters, especially the desired isoamyl acetate [48].
Author Copy
begin to run those metabolic pathways essential for cell growth [42]. When the
final objective of fermentation is both to achieve the highest levels of biomass
production and the optimal extraction of primary metabolites, what is sought
is to extend the growth phase, regularly by means of batch or continuous fed
cultures. Consequently, they begin with the stationary phase, in which fungal
biomass manages to remain constant and the growth rate returns to zero [41].
After prolonged periods in the stationary phase, individual cells can die and
self-autolyze [48]. The stationary phase is characterized by being the stage in
which the microorganism reaches long periods of survival without the need
for the addition of nutrients. Not only the absence of nutrients can induce the
stationary phase, there are other facilitating causes such as the presence of
toxic metabolites (for example, ethanol in the case of yeast), low pH, high
CO2, variable O2 and high temperature [30]. During the stationary phase of
unbalanced growth, fungi can undergo secondary metabolism, specifically
initiating metabolic pathways that are not essential for cell growth but are
involved in the organism’s survival. Therefore, the industrial production
of secondary fungal metabolic compounds such as penicillin and ergot
alkaloids essentially compromise the stability of the cell population during
the stationary phase [30, 41].
There are three processes that are commonly used in the production
of bioethanol which are separate hydrolysis and fermentation (SHF),
simultaneous saccharification and fermentation (SSF), and simultaneous
saccharification and co-fermentation (SSCF). In SHF, the hydrolysis of
lignocellulosic materials is separated from ethanol fermentation, in which
the separation of enzymatic hydrolysis and fermentation makes it easier for
the enzyme to operate at high temperature for better performance, whereas
fermentation organisms can be operated at a moderate temperature to
optimize the use of sugar. In contrast, the other two methods (SSF and SSCF)
have a short general process since the enzymatic hydrolysis and fermentation
process occurs simultaneously to keep the glucose concentration low; that is,
for SSF, the fermentation of glucose is separated from the pentose while the
SSCF ferments the glucose and the pentose in the same reactor. In turn, these
Apple Academic Press
processes offer a series of benefits such as lower cost, higher ethanol yield
and less processing time [49–52].
Bioethanol fermentation can be carried out in batches, batch feeds, repeat
batches or in continuous mode. In batch process, the substrate is provided at
the beginning of the process without adding or removing the medium; this
system is known as simpler bioreactor with multi-vessel control process, flex-
ible, and easy; where, the fermentation process is carried out in a closed-circuit
Author Copy
system with a high concentration of sugars and inhibitors at the beginning
and ends with a high concentration of product. Therefore, the batch system
offers a number of benefits, which include complete sterilization, requires
no job skills, is easy to handle raw materials, can be easily controlled, and
flexible to various product specifications. However, productivity is low and
requires high and intensive labor costs. The presence of a high concentration
of sugar in the fermentation medium can lead to inhibition of the substrate
and inhibit cell growth and the production of ethanol [1, 11, 41].
Cellular Recycling Batch Fermentation is a strategic method for effec-
tive ethanol production, as it reduces the time and cost of inoculum prepara-
tion; some of its advantages are easy cell harvesting, stable operation, and
long-term productivity. Besides, sugar materials and immobilized yeast
cells are used as facilitators in the cell separation necessary to carry out cell
recycling [37, 48].
Author Copy
gasoline in common spark-ignition engines without further modification or
can be operated at higher percentages in mixtures with gasoline in slightly
modified engines which are called flexi-fuel vehicles. It is possible to use
pure gasoline or ethanol blends up to 85% (even 100% ethanol in Brazil)
in flex-fuel engines [58].
The manufacture of these biofuels has increased since 2000 [60]. This
being the case, ethanol is more significant when it is considered according
to the volume produced. Today, the United States and Brazil are the world’s
largest ethanol producers, able to generate ethanol production exceeding 94
billion liters per year, which represents around 85% of world production
[61]. In particular, Brazil is the leading manufacturer of ethanol for auto-
mobiles, with an annual production rate of 4 billion gallons of ethanol from
sugar [62].
In 2016, biofuels supplied around 4.5% of total fuel for road transpor-
tation worldwide and its planned share of world transport fuel by 2050 is
estimated at 25%. [53]. As the lowest-cost manufacturer globally, the United
States remains to retain its place as the most reliable and affordable source of
ethanol internationally, reaching up to 57.8 billion liters of global manufacture
in 2016. Brazil, which produced roughly 27.6 billion L, is accountable for
about 27% of world production, while the EU follows with 5%. Other proper
leaders to mention are China and Canada about ethanol manufacture [57].
The geographic distribution of the production and consumption of ethanol
is related to many factors, such as manufacturing destinations, government
policies, natural resources availability, and environmental regulations.
Different world regions can be perceived as distinct markets with diverse
demands and supply. Estimates denote that the USA is clearly the largest
manufacturer and consumer of ethanol and it is followed by Brazil [63]. With
regards to assessing these main producers, a difference from the perspective
of the increase in production by 2020 is noted [59].
Author Copy
most important as input in ethanol production processes. Even so, under
the correct conditions of both feeding and environment, it is currently
possible to obtain volumes of more than 20% of ethanol production, when
these conditions are not met, leading to stress in the lead, stuck, slow, and
inefficient fermentations are obtained. Therefore, it is essential to optimize
alcoholic fermentations to understand aspects of yeast cell physiology,
particularly when lignocellulosic substrates are used for the production of
second-generation bioethanol.
KEYWORDS
• alcohol dehydrogenase
• bioethanol production
• biorefinery
• growth conditions
• lignocellulosic biomass
• yeast physiology
REFERENCES
1. Dzialo, M. C., Park, R., Steensels, J., Lievens, B., & Verstrepen, K. J., (2017). Physiology,
ecology, and industrial applications of aroma formation in yeast. FEMS Microbiol. Rev.,
41, S95–S128. https://doi.org/10.1093/femsre/fux031.
2. Buzzini, P., (2006). Yeast biodiversity and biotechnology. Yeast Handbook; Biodivers.
Ecophysiol. Yeasts, pp. 533–559. https://doi.org/10.1007/3-540-30985-3_22.
3. Van, D. S. P., (2015). Approaches to production of natural flavors. In: Parker, J. K.,
Elmore, J. S., & Methven, L. B., (eds.), Flavor Development, Analysis and Perception
in Food and Beverages (pp. 235–248). Woodhead Publishing. https://doi.org/https://doi.
org/10.1016/B978-1-78242-103-0.00011-4.
Apple Academic Press
4. Araújo, W. A., (2016). Ethanol industry: Surpassing uncertainties and looking forward.
In: Global Bioethanol. (pp. 1–33). Academic Press.
5. Bhatia, L., Johri, S., & Ahmad, R., (2012). An economic and ecological perspective of
ethanol production from renewable agro waste: A review. AMB Express, 2(1), 1–19.
6. Câmara, M. M., Soares, R. M., Feital, T., Naomi, P., Oki, S., Thevelein, J. M., & Pinto,
J. C., (2017). On-line identification of fermentation processes for ethanol production.
Bioprocess and Biosystems Engineering, 40(7), 989–1006.
7. Lopes, M. L., De Lima, P. S. C., Godoy, A., Cherubin, R. A., Lorenzi, M. S., Giometti,
F. H. C., & De Amorim, H. V., (2016). Ethanol production in Brazil: A bridge between
Author Copy
science and industry. Brazilian Journal of Microbiology, (47), 64–76.
8. Serafim, F. A. T., & Lanças, F. M., (2019). Sugarcane spirits (cachaça) quality assurance
and traceability: An analytical perspective. In: Grumezescu, A. M., & Holban, A. M.,
(eds.), Production and Management of Beverages (pp. 335–359). Elsevier Inc. https://
doi.org/10.1016/B978-0-12-815260-7.00011-0.
9. Vasconcelos De, J. N., (2015). Ethanol fermentation. In: Santos, F., Caldas, C., & Borém,
A., (eds.), Sugarcane: Agricultural Production, Bioenergy, and Ethanol (pp. 311–340).
Elsevier Inc. All. https://doi.org/10.1016/B978-0-12-802239-9.00015-3.
10. Dombek, K. M., & Ingram, L., (1987). Ethanol production during batch fermentation
with saccharomyces cerevisiae: Changes in glycolytic enzymes and internal PH. Appl.
Environ. Microbiol., 53, 1286–1291.
11. Sarris, D., & Papanikolaou, S., (2016). Biotechnological production of ethanol:
Biochemistry, processes and technologies. Eng. Life Sci., 16(4), 307–329. https://doi.
org/10.1002/elsc.201400199.
12. Zhang, B., Li, L., Zhang, J., & Gao, X., (2013). Improving ethanol and xylitol fermentation at
elevated temperature through substitution of xylose reductase in Kluyveromyces marxianus.
J. Ind. Microbiol. Biotechnol., 40, 305–316. https://doi.org/10.1007/s10295-013-1230-5.
13. Arshadi, M., & Grundberg, H., (2011). Biochemical production of bioethanol. In: Luque,
R., Campelo, J., & Clark, J., (eds.), Handbook of Biofuels Production (pp. 199–220).
Woodhead Publishing Limited. https://doi.org/10.1533/9780857090492.2.199.
14. Alperstein, L., Gardner, J. M., Sundstrom, J. F., Sumby, K. M., & Jiranek, V., (2020).
Yeast bioprospecting versus synthetic biology — which is better for innovative beverage
fermentation? Appl. Microbiol. Biotechnol., 104, 1939–1953. https://doi.org/10.1007/
s00253-020-10364-x.
15. Akhtar, N., Karnwal, A., Upadhyay, A. K., Paul, S., & Mannan, M. A., (2018).
Saccharomyces cerevisiae bio-ethanol production, a sustainable energy alternative.
Article Asian J. Microbiol. Biotechnol. Environ. Sci., 20, 1–5.
16. Hong, K. K., & Nielsen, J., (2012). Metabolic engineering of Saccharomyces cerevisiae:
A key cell factory platform for future biorefineries. Cell. Mol. Life Sci., 69, 2671–2690.
https://doi.org/10.1007/s00018-012-0945-1.
17. Techaparin, A., Thanonkeo, P., & Klanrit, P., (2017). Biotechnology and industrial
microbiology high-temperature ethanol production using thermotolerant yeast newly
isolated from greater Mekong subregion. Brazilian J. Microbiol., 48, 461–475. https://
doi.org/10.1016/j.bjm.2017.01.006.
18. Joshi, J., Dhungana, P., Prajapati, B., Maharjan, R., Poudyal, P., Yadav, M., Mainali, M., et
al., (2019). Enhancement of ethanol production in electrochemical cell by saccharomyces
cerevisiae (CDBT2) and wickerhamomyces anomalus. Front. Energy Res., 7, 1–11.
https://doi.org/10.3389/fenrg.2019.00070.
Apple Academic Press
19. Sukwong, P., Sunwoo, I. Y., Lee, M. J., & Ra, C. H., (2018). Application of the severity
factor and HMF removal of red macroalgae Gracilaria verrucosa to production of
bioethanol by Pichia stipitis and Kluyveromyces marxianus with adaptive evolution.
Appl. Biochem. Biotechnol.
20. Gargalo, C. L., Chairakwongsa, S., Quaglia, A., Sin, G., & Gani, R., (2015). Methods
and tools for sustainable chemical process design. In: Klemeš, J. J., (ed.), Assessing and
Measuring Environmental Impact and Sustainability (pp. 277–321). Elsevier Inc.
21. Quintero, J. A., Rincón, L. E., & Cardona, C. A., (2011). Production of bioethanol from
agro-industrial residues as feedstocks. In: Pandey, A., Ricke, S. C., Gnansounou, E.,
Author Copy
Larroche, C., & Dussap, C. G., (eds.), Biofuels: Alternative Feedstocks and Conversion
Processes (pp. 251–285). Elsevier Inc. https://doi.org/10.1016/C2010-0-65927-X.
22. Thammasittirong, S. N., Thirasaktana, T., Thammasittirong, A., & Srisodsuk, M., (2013).
Improvement of ethanol production by ethanol-tolerant Saccharomyces cerevisiae
UVNR56. SpringerPlus, 2, 1–5. https://doi.org/10.1186/2193-1801-2-583.
23. Watanabe, T., Watanabe, I., Yamamoto, M., Ando, A., & Nakamura, T., (2011). Bioresource
technology a UV-Induced mutant of Pichia stipitis with increased ethanol production
from xylose and selection of a spontaneous mutant with increased ethanol tolerance.
Bioresour. Technol., 102(2), 1844–1848. https://doi.org/10.1016/j.biortech.2010.09.087.
24. Arshad, M., Hussain, T., Iqbal, M., & Abbas, M., (2017). Enhanced ethanol production at
commercial scale from molasses using high gravity technology by mutant S. cerevisiae.
Brazilian J. Microbiol., 48, 403–409. https://doi.org/10.1016/j.bjm.2017.02.003.
25. Chen, H., & Fu, X., (2016). industrial technologies for bioethanol production from
lignocellulosic biomass. Renew. Sustain. Energy Rev., 57, 468–478. https://doi.org/10.1016/
j.rser.2015.12.069.
26. OECD/FAO, (2017). “Biofuels,” in OECD-FAO Agricultural Outlook, 2017–2026,
OECD Publishing, París. https://doi.org/10.1787/agr_outlook-2017-en.
27. Mohd, A. S. H., Abdulla, R., Jambo, S. A., Marbawi, H., Gansau, J. A., Mohd, F. A.
A., & Rodrigues, K. F., (2017). Yeasts in sustainable bioethanol production: A review.
Biochem. Biophys. Reports, 10, 52–61. https://doi.org/10.1016/j.bbrep.2017.03.003.
28. Yang, B., Dai, Z., Ding, S. Y., & Wyman, C. E., (2011). Enzymatic hydrolysis of
cellulosic biomass. Biofuels, 2(4), 421–449. https://doi.org/10.4155/bfs.11.116.
29. Olsson, L., & Hahn-Hägerdal, B., (1996). Fermentation of lignocellulosic hydrolysates for
ethanol production. Enzyme Microb. Technol., 18(5), 312–331. https://doi.org/10.1016/
0141-0229(95)00157-3.
30. Rehman, A., Tong, Q., Jafari, S. M., Assadpour, E., Shehzad, Q., Aadil, R. M., Iqbal,
M. W., et al., (2020). Carotenoid-loaded nanocarriers: A comprehensive review. Adv.
Colloid Interface Sci., 275, 102048. https://doi.org/10.1016/j.cis.2019.102048.
31. Walker, G. M., & White, N. A., (2005). Introduction to fungal physiology. Fungi Biol.
Appl., 1–34. https://doi.org/10.1002/0470015330.ch1.
32. Sun, Y., & Cheng, J., (2002). Hydrolysis of lignocellulosic materials for ethanol
production: A review. Bioresour. Technol., 83(1), 1–11. https://doi.org/10.1016/S0960-
8524(01) 00212-7.
33. Jin, H., Liu, R., & He, Y., (2012). Kinetics of batch fermentations for ethanol production
with immobilized Saccharomyces cerevisiae growing on sweet sorghum stalk juice.
Procedia Environ. Sci., 12, 137–145. https://doi.org/10.1016/j.proenv.2012.01.258.
34. Rubio-Arroyo, M. F., Vivanco-Loyo, P., Juárez, M., Poisot, M., & Ramírez-Galicia, G.,
Apple Academic Press
Author Copy
(Vol. 105). Elsevier Ltd. https://doi.org/10.1016/bs.aambs.2018.05.003.
39. Castaño, H., & Mejia, C., (2008). Production of ethanol from cassava starch using the
simultaneous saccharification-fermentation process strategy (SSF). Vitae, Rev. La Fac.
Química Farm., 15(2), 251–258.
40. Kostas, E. T., White, D. A., Du, C., & Cook, D. J., (2016). Selection of yeast strains for
bioethanol production from UK seaweeds. J. Appl. Phycol., 28(2), 1427–1441. https://
doi.org/10.1007/s10811-015-0633-2.
41. Aksu, Z., & Eren, A. T., (2007). Production of carotenoids by the isolated yeast of
Rhodotorula glutinis. Biochem. Eng. J., 35(2), 107–113. https://doi.org/10.1016/j.bej.
2007.01.004.
42. Walker, G. M., & Basso, T. O., (2020). Mitigating stress in industrial yeasts. Fungal
Biol., 124(5), 387–397. https://doi.org/10.1016/j.funbio.2019.10.010.
43. Saini, P., Beniwal, A., Kokkiligadda, A., & Vij, S., (2018). Response and tolerance of
yeast to changing environmental stress during ethanol fermentation. Process Biochem.,
72, 1–12. https://doi.org/10.1016/j.procbio.2018.07.001.
44. Święciło, A., (2016). Cross-stress resistance in Saccharomyces cerevisiae yeast—new
insight into an old phenomenon. Cell Stress Chaperones, 21(2), 187–200. https://doi.
org/10.1007/s12192-016-0667-7.
45. Van, D. M., Erdei, B., Galbe, M., Nygård, Y., & Olsson, L., (2019). Strain-dependent variance
in short-term adaptation effects of two xylose-fermenting strains of Saccharomyces cere-
visiae. Bioresour. Technol., 292, 121922. https://doi.org/10.1016/j.biortech.2019.121922.
46. Du Preez, J., (1994). Process parameters and environmenta1 factors affecting D-xylose
fermentation by yeasts. Enzyme Microb. Technol., 16, 944–956.
47. Timoumi, A., Guillouet, S. E., Molina-Jouve, C., Fillaudeau, L., & Gorret, N., (2018).
Impacts of environmental conditions on product formation and morphology of yarrowia
lipolytica. Appl. Microbiol. Biotechnol., 102(9), 3831–3848. https://doi.org/10.1007/
s00253-018-8870-3.
48. Zemančíková, J., Kodedová, M., Papoušková, K., & Sychrová, H., (2018). Four
saccharomyces species differ in their tolerance to various stresses though they have
similar basic physiological parameters. Folia Microbiol. (Praha), 63(2), 217–227.
https://doi.org/10.1007/s12223-017-0559-y.
49. Kucharczyk, K., & Tuszyński, T., (2017). The effect of wort aeration on fermentation,
maturation and volatile components of beer produced on an industrial scale. J. Inst.
Brew., 123(1), 31–38. https://doi.org/10.1002/jib.392.
50. Ishizaki, H., & Hasumi, K., (2013). Ethanol Production from Biomass. Elsevier, https://
doi.org/10.1016/B978-0-12-404609-2.00010-6.
51. Ravanal, M. C., Camus, C., Buschmann, A. H., Gimpel, J., Olivera-Nappa, Á., Salazar,
O., & Lienqueo, M. E., (2019). Production of bioethanol from brown algae. Adv. Feed.
Apple Academic Press
Convers. Technol. Altern. Fuels Bioprod. New Technol. Challenges Oppor., 69–88. https://
doi.org/10.1016/B978-0-12-817937-6.00004-7.
52. Teter, S. A., Sutton, K. B., & Emme, B., (2014). Enzymatic Processes and Enzyme
Development in Biorefining. https://doi.org/10.1533/9780857097385.1.199.
53. Silveira, M. H. L., Vanelli, B. A., & Chandel, A. K., (2017). Second Generation Ethanol
Production: Potential Biomass Feedstock, Biomass Deconstruction, and Chemical
Platforms for Process Valorization. Potential biomass feedstock, biomass deconstruction,
and chemical platforms for process valorization. Elsevier Inc. https://doi.org/10.1016/
B978-0-12-804534-3.00006-9.
Author Copy
54. International Energy Agency (IEA), (2011). Technology Roadmap: Biofuels for
Transport. Organization for Economic Cooperation and Development, IEA, Paris.
http://www.iea.org/publications/freepublications/publication/biofuels_roadmap.pdf
(accessed on 28 October 2021).
55. Joshi, V. K., Walia, A., & Rana, N. S., (2012). Production of bioethanol from food
industry waste: Microbiology, biochemistry, and technology. In: Biomass Conversion
(pp. 251–311). Springer, Berlin, Heidelberg.
56. Ghosh, T. K., & Prelas, M. A., (2011). Ethanol. In: Energy Resources and Systems (pp.
419–493). Springer, Dordrecht.
57. Matsuoka, S., Ferro, J., & Arruda, P., (2009). The Brazilian experience of sugarcane
ethanol industry. In Vitro Cellular & Developmental Biology-Plant, 45(3), 372–381.
58. Kohler, M., (2019). Economic assessment of ethanol production. In: Ethanol, (pp.
505–521) Elsevier.
59. Amiri, T. Y., & Ghasemzadeh, K., (2019). Ethanol economy: Environment, demand, and
marketing. In: Ethanol. (451–504). Elsevier.
60. Kolling, D. F., Dalla, C. V. F., & Oliveira, C. A. O., (2014). Global market issues in the
liquid biofuels industry. In: Liquid Biofuels: Emergence, Development and Prospects
(pp. 55–72). Springer, London.
61. Lopes, M. L., Cristina, S., Paulillo, D. L., Godoy, A., Cherubin, R. A., Lorenzi, M. S.,
Henrique, F., et al., (2016). Ethanol production in Brazil: A bridge between science and
industry. Brazilian J. Microbiol., 1–13. https://doi.org/10.1016/j.bjm.2016.10.003.
62. Anđelković, D., Antić, B., Vujanić, M., Subotić, M., & Radovanović, L., (2017). The
perspectives of applying ethanol as an alternate fuel. Energy Sources, Part B Econ.
Planning, Policy, 12(9), 1–10. https://doi.org/10.1080/15567249.2012.683930.
63. Janda, K., Kristoufek, L., & Zilberman, D., (2012). Biofuels: Policies and impacts.
Agricultural Economics, 58(8), 372–386.
Author Copy
NORMA M. DE LA FUENTE-SALCIDO
Faculty of Biological Sciences, Autonomous University of Coahuila,
Torreón, Coahuila, México, E-mail: normapbr322@gmail.com
(N. M. D. L. Fuente-Salcido)
ABSTRACT
2.1 INTRODUCTION
Since ancient civilization, the use of alcoholic beverages has taken part of the
universal culture, traditions, and the economic strength of nations. However,
that is why they thought that initially fermentations were considered a
importance, and research. For this process, there are different sources of
sugars such as cereals, fruits, and other carbon sources that can be used by
various microorganisms to obtain products through the fermentation process
such as wine, beer, or even biofuel that are economically essential products
(Figure 2.1).
Author Copy
FIGURE 2.1 The fermentation process for the elaboration of alcoholic beverages and biofuel
with different carbon sources.
Author Copy
to the yield, the factors mentioned previously can change depending on
the microorganism used, as well as the type of fermentation that is carried
out, and the parameters that are used to obtain the highest production of
ethanol per gram of carbon source. In this chapter aims to integrate the
current knowledge about the differences in the yeast S. cerevisiae and the
bacterium Z. mobilis regarding to ethanolic fermentation. Also, the central
role, the potential use of biological features of Z. mobilis and the common
biotechnological mechanisms and tools for the production of bioethanol
will be discussed.
Ancient civilizations brewed beer around the year 6000 A.C. through the
fermentation process. For 2017, the average ethanol production is more
than 1 million barrels (159 million liters) per day, with an annual rate of
16 million gallons (60 million liters). There are many microorganisms that
produce ethanol but in each one has certain limitations due to economic
problems of the production of bioethanol, such as the use of different
substrates and yield. Within the microorganisms, yeasts are leaders in the
production of bioethanol, however, bacteria such as Z. mobilis, Escherichia
coli, or Clostridium sp., are being developed to increase the production
of bioethanol and eliminate the limitations such as the kind of substrate
or tolerance to different factors that can change with the help of genetic
engineering [5].
In addition, ethanol derived from yeast fermentation represents more
than 80% of the world’s renewable fuels and more than 85% of the alcohol
produced is provided by the United States and Brazil. The S. cerevisiae
produces more than 100 billion liters of ethanol per year [6, 7].
Author Copy
between the inside and the outside of the cell [9].
Both Z. mobilis and S. cerevisiae have a different pathway to use a carbon
source during the bioethanol production. Alcoholic fermentation commonly
includes diverse biochemical degradation pathways as glycolysis, alcoholic
fermentation, glycerol-pyruvic fermentation. Also, respiration processes
for utilization of hexoses, xylose catabolic pathways for utilization of
pentoses and glycerol assimilation, and glycolysis for glycerol-converting
microorganisms, and regulation between fermentation and respiration. In
the particular case of Z. mobilis uses ED pathway to anaerobically ferment
glucose for ethanol production. But S. cerevisiae uses EMP pathway for
glycolysis [2]. Fermentation process consists for Z. mobilis 1 mole of ATP is
yielded per mole of glucose through the ED pathway required less enzymatic
protein because together with Pdc and two alcohol dehydrogenases (Adh)
“form the backbone” of this glycolysis metabolic pathway. On the other
hand, S. cerevisiae one molecule of glucose (C6H12O6) is converted into two
molecules of pyruvic acid (C3H4O3) during the process of glycolysis. Pyruvic
acid is further decarboxylated to generate two molecules of acetaldehyde
(CH3CHO), which is reduced to ethanol (C2H5OH). The process described
produce two molecules of ATP as gain, two molecules of ethanol and two
molecules of CO2 [10]. In the conversion of glucose to ethanol and CO2
during fermentation, a total of 12 yeast enzymes are involved, 10 to degrade
glucose to pyruvate with the generation of ATP. After that, pyruvic acid is
converted to acetaldehyde through PDC and the final stage ADH converts
acetaldehyde to ethanol in this way NAD+ is regenerated to allow glycolysis
to continue (Figure 2.2) [7]. This has a big importance because Zymomonas
use less energy than Saccharomyces to make the fermentation process and
these can be traduced to “higher production of ethanol, using less energy to
make the process” depending on the strains.
Author Copy
FIGURE 2.2 Glycolysis and fermentation in yeast.
2.3.1 SUBSTRATES
2.3.2 MICROORGANISMS
and Rhizopus mobilis are capable to produce ethanol [11, 15]. However, not
all microorganisms can produce the same amount of ethanol, because each
one has their own limitations, like the type of fermentation or the different
tolerance like high sugar concentration or ethanol tolerance.
Apple Academic Press
2.3.3 PROCESS
Author Copy
found that make it possible to carry out the viable hydrolysis and fermenta-
tion process, the most common include separate hydrolysis and fermentation
(SHF), simultaneous saccharification and fermentation (SSF), simultaneous
saccharification and fermentation (SSCF), consolidated bioprocessing
(CBP), pre-saccharification followed by simultaneous saccharification and
fermentation (PSSF), separate hydrolysis and co-fermentation (SHCF), and
finally, pre-saccharification followed by simultaneous saccharification and
fermentation (PSSF) [16, 17].
2.3.4 YIELD
The yield depends on the amount of metabolite required, the amount produced
by the native organism and the biological activity of the compound [18].
Author Copy
isolates, a popular Mexican fermented pre-Hispanic beverage and reported
as “Thermobacterium mobile.” However, until 1936 Taxon Z. mobilis subsp.
mobilis according to the proposal of Kluyver and van Niel, and formalized in
1976 by the classification proposal and reported by De Ley, J., and J. Swings.
The commonly known taxonomic hierarchy of Zymomonas is described
as follows:
Phylum Proteobacteria
Class Alphaproteobacteria
Order Sphingomonadales
Family Sphingomonadaceae
Genus Zymomonas
Species Zymomonas mobilis
Subspecies Zymomonas mobilis francensis
Zymomonas mobilis mobilis
Zymomonas mobilis pomaceae
The bacteria description included in the Bergey’s Manual of Systematic
Bacteriology listed the characteristics of the bacterium [21].
This facultative aerobic Z. mobilis is, gram-negative bacteria, a non-spore-
forming, rod-shaped grouped in pairs and size of 2–6 × 1.0–1.4 µm.
Commonly non-motile but some exceptions may be possessed one to four
polar flagella. Growth of Z. mobilis on standard medium agar [D-glucose (20
g L–1), yeast extract (5 g L–1)], grows forming glistening colonies, regularly
edged, white to cream-colored, 1–2 mm in diameter after 48 h at 30°C
incubation. Growth optimal conditions includes 30°C and at pH 3.5–7.5 and
vitamins (biotin, pantothenate) as micronutrients [70]. Their nutrition includes
fermentable sugars (glucose, fructose, sucrose), is chemoorganotrophic
and highly efficient ethanol producer through the ED pathway [22]. Shows
constitutive ethanol-tolerance (5%) and also is acid tolerance, but is sensible
Author Copy
The physiology of a microorganism is decisive key to ensure the synthesis of
several high added value products by advanced fermentation technologies.
Z. mobilis physiology is particularly fascinating, mostly by their capacity
to synthesize economically and sustainably important products such as
bioethanol among others biochemical as bionic acid, levan, sorbitol, among
others [5, 28].
In this sense, the ED metabolic pathway of Zymomonas mobilis strains is
considered a unique ethanologenic pathway due to its efficiency, character-
ized by a decrease in cellular ATP consumption, which promotes an increase
in the transformation of sugars into biomass, and even with better yields and
productivity of the bioethanol found in S. cerevisiae [23]. The ethanologenic
Zymomonas strains are recognized as the most efficient bacterial ethanol
producer, mainly due to that employ the ED pathway as a strictly step during
a fermentative process, in addition to having a physiology perfectly adapted
to industrial-scale bioethanol production. Sequential steps to glucose trans-
formation by the ED pathway, the glucose phosphate is catabolized into
2-keto-3-deoxy-6-phosphogluconic acid (KDPG) and then is cleaved by
KDPG aldolase to pyruvate and glyceraldehyde-3-phosphate (GAP). The
GAP is oxidized to pyruvate by glycolytic enzymes and ATP produced by
substrate-level phosphorylation. The next reaction is reduction of pyruvic
acid to ethanol and CO2. The overall reaction is:
C6H12O6 → C2H5OH
Glucose → 2 Ethanol + 2 CO2 + 1 ATP
As a catabolic main route exclusive of prokaryotes, the enzymatic divergence
of the ED pathway is it uses 6-phosphogluconate dehydratase and 2-keto-3
deoxyphosphogluconate aldolase to synthesize pyruvate from glucose. There-
fore, the pathway rises a yield of 1 ATP for each glucose molecule catabolized,
as well as 1 NADH and 1 NADPH. The obvious comparison with the EMP
pathway, indicates that glycolysis rises a yield of 2 ATP and 2 NADH for each
glucose molecule processed [71]. This physiological feature distinguishes Z.
mobilis as a profitable strain for industrial applications.
Apple Academic Press
Author Copy
microorganisms generate mixtures of metabolic products in addition to
alcohol. The disadvantage of the use of Zymomonas lies in its specificity
for glucose as the only fermentable sugar and its lacks the enzymes to
break down another carbon source, including the enzymes to degrade
carbohydrate polymers such as starch and cellulose [31], and this would
limit its biotechnological application.
Different strategies, including kinetic modeling and rational metabolic
engineering to understand how central metabolism of Zymomonas work, and
novel genome sequence are widely studied and applied to improve ethanol
synthesis. In addition, the techniques promoted by the application of synthetic
biology, metabolic engineering, and the ongoing world trend of industrial
bioethanol production from the use of agro-industrial waste by the metabolism
of Zymomonas mobilis has facilitated the development of genetically modified
bacteria for co-culturing in cellulose-rich media [14, 32, 72–75].
Author Copy
FIGURE 2.3 Metabolic pathway of Saccharomyces cerevisiae and Zymomonas mobilis.
Knowing the differences between the different metabolic routes that exist
between the bacteria Z. mobilis and the yeast S. cerevisiae. Lack of different
enzymes in the Zymomonas metabolic pathway is found as phosphofruc-
tokinase (Pfk) in the EMP pathway, phosphogluconate dehydrogenase
(Pgd) and transaldolase (Tal) in the PPP pathway, as well as 2-oxoglutarate
dehydrogenase complex (sucABCD) and malate dehydrogenase (Mdh)
in the TCA cycle and this enzyme deficiency carry more carbon into the
ethanol production pathway and highly efficient glycolysis resulting in the
Author Copy
like 11–16% v/v [29].
Adh synthesizes the protein that transforms acetaldehyde into ethanol in the
metabolic pathway ED, however, genetic engineering is looking for different
ways to increase the production of ethanol and inhibit acetate production. On
the other hand, there are reports of strains that are already modified with the
Apple Academic Press
inclusion of different gene such as XylA, XylB, talB, and tktA that give them
the ability to use 5 C sugars and thus, be able to use expanding the range of
raw material used as a carbon source [38].
Regarding to S. cerevisiae have only 5,538 genes encoding 100 amino
acids and contains 18 genes for which orthologs are not identified. These
may be species-specific genes in S. cerevisiae, but alternatively, they could
reflect the gaps in the draft genomic sequences available [39, 40]. Yeast cells
Author Copy
exposed to ethanol synthesize a range of Hsps, including Hsp104, Hsp82,
Hsp70, Hsp26, Hsp30, and Hsp12. The Hsp104 and Hsp12 physiologically
influence the ethanol tolerance in yeast [40]. Cell wall from S. cerevisiae
contains 85% polysaccharides and 15% proteins approximately. If there is
an increase in the level of ethanol, the stability of the membrane will be
affected, which will cause damage to the proteins and, therefore, endocytosis
is inhibited through the membrane [10].
In Figure 2.4, the yeast the Hog1 gene is responsible for encoding the
high osmolarity glycerol pathway (HOG), this pathway mainly encodes two
enzymes GPD1 and GPD2 that catalyze the conversion of dihydroxyacetone
phosphate (DHAP) through glycerol-3-phosphate (G3P) to glycerol. The
Fps1 channel helps to glycerol accumulation and contributes to raising the
osmotic pressure within the cell. Within the stress response are the proteins
trehalose-6-phosphate synthase (TPS) and trehalose-6-phosphate phospha-
tase (TPP), enzymes responsible for the synthesis of trehalose. In the same
way, heat shock transcription factor (Hsf1) induces the production of HSPs
[40]. Also, in Figure 2.4 in Z. mobilis the Zms4 and Zms6 genes are involved
in direct sRNA and target mRNA interactions. The accumulation of ethanol
within cells under stress increases the expressions of these genes. Increasing
the level of Zms4 accelerates ethanol catabolism through upregulation of
the aldehyde dehydrogenase gene (SsdA/ZmsZMO1754) directly and the
alcohol dehydrogenase 1 gene (AdhA) indirectly to mask ethanol to other
carboxylic acids. On the other hand, Zms6 upregulated under ethanol stress
regulates the expression of the lysine export Zms1437 gene and negatively
regulates the expression of the methylase gene of Zms1934 N-6 DNA to
improve ethanol tolerance and avoid import of methylated DNA created by
ethanol damage, respectively. However, several regulatory functions of this
Zms6 gene are still unknown. Otherwise, Zms16 gene interacts directly as a
target only with the gen Zms6 [41].
Author Copy
FIGURE 2.4 Metabolic pathway of Saccharomyces cerevisiae and Zymomonas mobilis with
principal genes.
hours of fermentation, using Z. mobilis (ATCC 10988) strain, which uses the
ED fermentation pathway, and has a larger surface area than Saccharomyces.
Fatty acids are produced by yeast and bacteria in response to stress. In this
study, Wang et al. in 2016 [45] modified an operon (CP4) that induced their
overproduction and high ethanol yield of 78 gL–1 at 32°C in 60 h. Davis
in 2006 [42] evaluated the ethanol production by Z. mobilis ZM4 and, the
comparison of S. cerevisiae in the same conditions, using the same substrate.
Author Copy
Ethanol production was 0.49 gg–1 with 100 g of glucose at 30°C with Z.
mobilis ZM4 strain and 0.46 gg–1 with S. cerevisiae, this can be attributed
to the medium supplementing with KH2PO4 and (NH4) 2SO4 among other
substances [5, 42]. Mazaheri and Pirouzi in 2020 [46], evaluate Z. mobilis
for bioethanol production from potato peel waste, obtaining 23.3 gL–1 with
initial sugar concentration of 61.3 gL–1. The combination of enzymes for the
hydrolysis process could have effectively released the fermentable sugars
from the solids of the potato peel, however, the final concentration of ethanol
obtained is not high enough for industrial scales as it causes high energy
consumption in the purification stage [46].
The four major stress responses studied during fermentation are tempera-
ture stress, osmotic stress, lack of nutrients and ethanol produced during
fermentation [40].
There are different strains to Zymomonas that can support high concentration
of glucose to grow up in 400 gL–1 medium, but the optimal glucose concentra-
tion is 200 gL–1 to obtain the maximum concentration of ethanol, 96 gL–1. This
bacterium has a higher glucose tolerance than S. cerevisiae and, therefore,
highly concentrated glucose media can potentially be used. The bacterium
metabolizes glucose, sucrose, and fructose. However, there are some strains
that are modified genetically to use other carbon sources like xylose [47].
Author Copy
ethanol tolerance in yeast, as well as available nutrients and growth substrates
[50]. High ethanol tolerance is a hallmark of Saccharomyces cerevisiae and a
wide variety of genes are responsible for a moderate ethanol tolerance ranging
from 6 to 12% in laboratory strains and 16 to 20% in industrial strains, respec-
tively [79]. According to Ming [80], wild-type Z. mobilis has an adaptive
mechanism in response to ethanol stress and could tolerate up to 13%. There-
fore, with these results, we can see that Zymomonas has a great advantage
compared to yeast for tolerance to ethanol in the medium, this being one of
the most important factors to consider choosing a microorganism to carry out
the fermentation process in this bioprocess for ethanol production [81].
Temperature is one of the most important factors that affect the production
of ethanol by different microorganisms. The fermentation without cooling,
produces microbial heat shock stress, the microbial growth is reduced, and
ethanol production yield is affected. There are some Zymomonas strains
thermotolerant, can growth even at 39°C, that is, 5–10°C higher than the
optimum temperature. This Z. mobilis harbor degP gene from heat shock
proteins (HSP) family [51, 52]. The HSF family of proteins (heat shock
transcription factor) are primary modulators of heat shock response (HSR),
as well as the MSN2 and MSN4 genes of significant use in the expression of
the heat shock gene. Heat shock factor in S. cerevisiae (Hsf1) is an essential
protein that assists HSR in adapting to oxidative stress and glucose defi-
ciency [40].
It is noted that despite the fact that some Saccharomyces strains can grow
at high temperatures, ethanol production is lower than Zymomonas strains
that grow at the same range of temperatures [43, 52–56].
2.11 CONCLUSION
Author Copy
strongly suggest that the unique and amazing physiology of Z. mobilis provides
efficient biosynthetic machinery for industrial-scale biotechnological produc-
tion of bioethanol.
KEYWORDS
• bioethanol
• biofuels
• heat shock proteins
• Saccharomyces cerevisiae
• trehalose-6-phosphate phosphatase
• Zymomona mobilis
REFERENCES
1. Varela, C., (2016). The impact of non-Saccharomyces yeasts in the production of alcoholic
beverages. Applied Microbiology and Biotechnology, 100(23), 9861–9874. https://doi.
org/10.1007/s00253-016-7941-6.
2. Sarris, D., & Papanikolaou, S., (2016). Biotechnological production of ethanol:
Biochemistry, processes and technologies. Engineering in Life Sciences, 16(4), 307–329.
https://doi.org/10.1002/elsc.201400199.
3. Mohd, A. S. H., Abdulla, R., Jambo, S. A., Marbawi, H., Gansau, J. A., Mohd, F.
A. A., & Rodrigues, K. F., (2017). Yeasts in sustainable bioethanol production: A
review. Biochemistry and Biophysics Reports, 10, 52–61. https://doi.org/10.1016/j.
bbrep.2017.03.003.
4. Bleoanca, I., & Bahrim, G., (2013). Overview on brewing yeast stress factors. Romanian
Biotechnological Letters, 18(5), 8559–8572.
5. Yang, S., Fei, Q., Zhang, Y., Contreras, L. M., Utturkar, S. M., Brown, S. D., Himmel,
M. E., & Zhang, M., (2016). Zymomonas mobilis as a model system for production of
biofuels and biochemicals. Microb. Biotechnol., 9, 699–717.
6. Shaw, A. J., Lam, F. H., Hamilton, M., Consiglio, A., MacEwen, K., Brevnova, E. E., &
Apple Academic Press
Author Copy
9. Zhang, J., Zhang, W., & Gu, Y., (2018). Enzyme-free isothermal target-recycled
amplification combined with page for direct detection of microRNA-21. Analytical
Biochemistry, 550, 117–122. https://doi.org/10.1016/j.ab.2018.04.024.
10. Mannan, M. A., Upadhyay, A. K., & Karnwal, A., (2017). Saccharomyces cerevisiae
bio-ethanol production a sustainable energy. Asian Journal of Microbiology, Biotechnology
and Environmental Sciences, 20.
11. Gupta, A., & Prakash, J., (2015). Sustainable bio-ethanol production from agro-residues: A
review. Renewable and Sustainable Energy Reviews, 41, 550–567. https://doi.org/10.1016/j.
rser.2014.08.032.
12. Aditiya, H. B., Mahlia, T. M. I., Chong, W. T., Nur, H., & Sebayang, A. H., (2016).
Second-generation bioethanol production: A critical review. Renewable and Sustainable
Energy Reviews, 66, 631–653. https://doi.org/10.1016/j.rser.2016.07.015.
13. Souza, C. J. A. D., Costa, D. A., Rodrigues, M. Q. R. B., Ancély, F., Lopes, M. R.,
Abrantes, A. B. P., & Fietto, L. G., (2012). Bioresource technology the influence of
presaccharification, fermentation temperature and yeast strain on ethanol production
from sugarcane bagasse. Bioresource Technology, 109, 63–69. https://doi.org/10.1016/j.
biortech.2012.01.024.
14. Carrillo-Nieves, D., Rostro, A. M. J., De La Cruz, Q. R., Ruiz, H. A., Iqbal, H. M. N.,
& Parra-Saldívar, R., (2019). Current status and future trends of bioethanol production
from agro-industrial wastes in Mexico. Renewable and Sustainable Energy Reviews,
102, 63–74. https://doi.org/10.1016/j.rser.2018.11.031.
15. Rastogi, M., & Shrivastava, S., (2017). Recent advances in second-generation bioethanol
production: An insight to pretreatment, saccharification and fermentation processes.
Renewable and Sustainable Energy Reviews, 80, 330–340. https://doi.org/10.1016/j.
rser.2017.05.225.
16. Carrillo-nieves, D., Rostro, M. J., De, R., & Quiroz, C., (2019). Current Status and
Future Trends of Bioethanol Production from Agro-Industrial Wastes in Mexico, 102,
63–74. https://doi.org/10.1016/j.rser.2018.11.031.
17. Jahnavi, G., Prashanthi, G. S., Sravanthi, K., & Rao, L. V., (2017). Status of availability
of lignocellulosic feedstocks in India: Biotechnological strategies involved in the
production of bioethanol. Renewable and Sustainable Energy Reviews, 73, 798–820.
https://doi.org/10.1016/j.rser.2017.02.018.
18. Connor, S. E. O., (2015). Engineering of Secondary Metabolism, 1–24. https://doi.
org/10.1146/annurev-genet-120213-092053.
19. Pappas, K. M., Kouvelis, V. N., Saunders, E., Brettin, T. S., Bruce, D., Detter, C.,
Balakireva, M., et al., (2011). Genome sequence of the ethanol-producing Zymomonas
mobilis subsp. mobilis lectotype strain ATCC 10988. Journal of Bacteriology, 193(18),
5051, 5052. https://doi.org/10.1128/JB.05395-11.
Apple Academic Press
20. Ojeda-Linares, C. I., Vallejo, M., Lappe-Oliveras, P., & Casas, A., (2020). Traditional
management of microorganisms in fermented beverages from cactus fruits in Mexico:
An ethnobiological approach. J. Ethnobiol. Ethnomed., 16(1), 1. doi: 10.1186/
s13002-019-0351-y.
21. Bergey’s Manual of Systematics of Archaea and Bacteria, Online © 2015 Bergey’s Manual
Trust. This article is © 2005 Bergey’s Manual Trust. doi: 10.1002/9781118960608.
gbm00925. Published by John Wiley & Sons, Inc., in association with Bergey’s Manual
Trust.
22. Chacon-Vargas, K., Chirino, A. A., Davis, M. M., Debler, S. A., Haimer, W. R., Wilbur, J.
Author Copy
J., Mo, X., et al., (2017). Genome sequence of Zymomonas mobilis subsp. mobilis NRRL
B-1960. Genome Announc., 5(30), e00562–17. https://doi.org/10.1128/genomeA.00562-17.
23. Geng, B., Cao, L., Li, F., Song, H., Liu, C. H., Zhao, X. Q., & Bai, F. W., (2020).
Potential of Zymomonas mobilis as an electricity producer in ethanol production.
Biotechnol Biofuels, 13, 36. https://doi.org/10.1186/s13068-020-01672-5.
24. Musatti, A., Cappa, C., Mapelli, C., Alamprese, C., & Rollini, M., (2020). Zymomonas
mobilis in bread dough: Characterization of dough leavening performance in presence
of sucrose. Foods, 9: 89. doi: 10.3390/foods9010089.
25. Wang, W., Dai, L., Wu, B., Bu-Fan, Q., Tian-Fang, H., Guo-Quan, H., & Ming-Xiong,
H., (2020). Biochar-mediated enhanced ethanol fermentation (BMEEF) in Zymomonas
mobilis under furfural and acetic acid stress. Biotechnol Biofuels, 13, 28. https://doi.
org/10.1186/s13068-020-1666-6.
26. Musatti, A., Mapelli, C., Rollini, M., Foschino, R., & Picozzi, C., (2018). Can Zymomonas
mobilis substitute Saccharomyces cerevisiae in cereal dough leavening? Foods, 7, 61.
27. Rogers, P. L., Jeon, Y. J., Lee, K. J., & Lawford, H. G., (2007). Zymomonas mobilis for
fuel ethanol and higher-value products. Adv Biochem Eng Biotechnol., 108, 263–288.
doi: 10.1007/10_2007_060.
28. Zhang, K., et al., (2019). New technologies provide more metabolic engineering
strategies for bioethanol production in Zymomonas mobilis. Applied Microbiology and
Biotechnology, 103, 2087–2099.
29. Fuchino, K., Kalnenieks, U., Rutkis, R., Grube, M., & Bruheim, P., (2020). Metabolic
Profiling of Glucose‐Fed Metabolically Active Resting Zymomonas Mobilis Strains
Metabolites, 10, 81. doi: 10.3390/metabo10030081.
30. Carreón-Rodríguez, O. E., Gutiérrez-Ríos, R. M., Acosta, J. L., Martinez, A., & Cevallos,
M. A., (2019). Phenotypic and genomic analysis of Zymomonas mobilis ZM4 mutants
with enhanced ethanol tolerance. Biotechnology Reports (Amsterdam, Netherlands), 23,
e00328. https://doi.org/10.1016/j.btre.2019.e00328.
31. Clark, & Pazdernik, (2016). Synthetic biology. In: Biotechnology. Elsevier. http://
dx.doi.org/10.1016/B978-0-12-385015-7.00013-2.
32. Fuchino, K., Kalnenieks, U., Rutkis, R., Grube, M., & Bruheim, P., (2020). Metabolic
profiling of glucose-fed metabolically active resting zymomonas mobilis strains.
Metabolites, 10(3), 6–8. https://doi.org/10.3390/metabo10030081.
33. Belda, I., Ruiz, J., Esteban-Fernández, A., Navascués, E., Marquina, D., Santos, A., &
Moreno-Arribas, M. V., (2017). Microbial contribution to wine aroma and its intended
org/10.1016/j.ymben.2018.04.001.
35. Lee, K. Y., Park, J. M., Kim, T. Y., Yun, H., & Lee, S. Y., (2010). The genome-scale
metabolic network analysis of Zymomonas mobilis ZM4 explains physiological features
and suggests ethanol and succinic acid production strategies. Microbial Cell Factories,
9, 1–12. https://doi.org/10.1186/1475-2859-9-94.
36. Annaluru, N., Annaluru, N., Muller, H., Mitchell, L. A., Ramalingam, S., Stracquadanio,
G., & Han, J. S., (2014). Total Synthesis of a Functional Designer Eukaryotic
Chromosome, 55. https://doi.org/10.1126/science.1249252.
37. Goffeau, A., Barrell, B. G., Bussey, H., Davis, R. W., Dujon, B., Feldmann, H., & Oliver,
Author Copy
S. G., (1996). Life with 6000 genes conveniently among the different interna- old
questions and new answers the genome. At the beginning of the se- of its more complex
relatives in the eukary- cerevisiae has been completely sequenced Schizosaccharomyces
pombe indicate. Science, 274, 546–567. https://doi.org/jyu.
38. He, M. X., Wu, B., Qin, H., Ruan, Z. Y., Tan, F. R., Wang, J. L., & Hu, Q. C., (2014).
Zymomonas mobilis: A novel platform for future biorefineries. Biotechnology for
Biofuels, 7(1), 1–15. https://doi.org/10.1186/1754-6834-7-101.
39. Kellis, M., Patterson, N., Endrizzi, M., Birren, B., & Lander, E. S., (2003). Sequencing
and comparison of yeast species to identify genes and regulatory elements. Nature,
423(6937), 241–254. https://doi.org/10.1038/nature01644.
40. Saini, P., Beniwal, A., Kokkiligadda, A., & Vij, S., (2018). Response and tolerance
of yeast to changing environmental stress during ethanol fermentation. Process
Biochemistry, 72, 1–12. https://doi.org/10.1016/j.procbio.2018.07.001.
41. Han, R., Haning, K., Gonzalez-rivera, J. C., Yang, Y., & Li, R., (2020). Multiple Small
RNAs Interact to Co-regulate Ethanol Tolerance in Zymomonas Mobilis, 8, 1–19.
https://doi.org/10.3389/fbioe.2020.00155.
42. Davis, L., Rogers, P., Pearce, J., & Peiris, P., (2006). Evaluation of Zymomonas -based
Ethanol Production from a Hydrolyzed Waste Starch Stream, 30, 809–814. https://doi.
org/10.1016/j.biombioe.2005.05.003.
43. Benjaphokee, S., Hasegawa, D., Yokota, D., Asvarak, T., Auesukaree, C., Sugiyama, M.,
& Harashima, S., (2012). Highly efficient bioethanol production by a Saccharomyces
cerevisiae strain with multiple stress tolerance to high temperature acid and ethanol.
New Biotechnology, 29(3), 379–386. https://doi.org/10.1016/j.nbt.2011.07.002.
44. Panesar, P. S., Marwaha, S. S., & Kennedy, J. F., (2006). Zymomonas Mobilis: An
Alternative Ethanol Producer, 635, 623–635. https://doi.org/10.1002/jctb.1448.
45. Wang, H., Cao, S., Tianshuo, W., Kaven, W., & Wang, T., (2016). Very high gravity
ethanol and fatty acid production of Zymomonas mobilis without amino acid and
vitamin. Journal of Industrial Microbiology & Biotechnology. https://doi.org/10.1007/
s10295-016-1761-7.
46. Mazaheri, D., & Pirouzi, A., (2020). Valorization of Zymomonas mobilis for bioethanol
production from potato peel: Fermentation process optimization. Biomass Conv. Bioref.
https://doi.org/10.1007/s13399-020-00834-7.
47. Ajit, A., Sulaiman, A. Z., & Chisti, Y., (2017). Production of bioethanol by Zymomonas
mobilis in high-gravity extractive fermentations. Food and Bioproducts Processing,
102, 123–135. https://doi.org/10.1016/j.fbp.2016.12.006.
48. Gonzalez, R., Quirós, M., & Morales, P. (2013). Yeast respiration of sugars by
non-Saccharomyces yeast species: a promising and barely explored approach to lowering
alcohol content of wines. Trends in Food Science & Technology, 29(1), 55–61. https://
doi.org/10.1016/j.tifs.2012.06.015.
Apple Academic Press
49. Boynton, P. J., & Greig, D., (2014). The Ecology and Evolution of Non-domesticated
Saccharomyces Species, 449–462. https://doi.org/10.1002/yea.
50. Riles, L., & Fay, J. C., (2019). Genetic basis of variation in heat and ethanol tolerance in
Saccharomyces cerevisiae. G3 (Bethesda, Md.), 9(1), 179–188. https://doi.org/10.1534/
g3.118.200566.
51. Anggarini, S., Murata, M., Kido, K., Kosaka, T., Sootsuwan, K., Thanonkeo, P., &
Yamada, M., (2020). Improvement of thermotolerance of Zymomonas mobilis by genes
for reactive oxygen species-scavenging enzymes and heat shock proteins. Frontiers in
Author Copy
Microbiology, 10, 1–14. https://doi.org/10.3389/fmicb.2019.03073.
52. Charoensuk, K., Sakurada, T., Tokiyama, A., Murata, M., Kosaka, T., Thanonkeo,
P., & Yamada, M., (2017). Thermotolerant genes essential for survival at a critical
high temperature in thermotolerant ethanologenic Zymomonas mobilis TISTR 548.
Biotechnology for Biofuels, 10(1), 1–11. https://doi.org/10.1186/s13068-017-0891-0.
53. D’Amore, T., Celotto, G., Russell, I., & Stewart, G. G., (1989). Selection and optimization
of yeast suitable for ethanol production at 40 C. Enzyme and Microbial Technology, 11(7),
411–416. https://doi.org/10.1016/0141-0229(89)90135-X.
54. Sree, N. K., Sridhar, M., Suresh, K., Banat, I. M., & Rao, L. V., (2000). Isolation of ther-
motolerant, osmotolerant, flocculating Saccharomyces cerevisiae for ethanol production.
Bioresource Technology, 72(1), 43–46. https://doi.org/10.1016/S0960-8524(99)90097-4.
55. De Souza, E. A., Gomes, D. S., Panek, A. D., & Eleutherio, E. C. A., (2003). The role
of glutathione in yeast dehydration tolerance. Cryobiology, 47(3), 236–241. https://doi.
org/10.1016/j.cryobiol.2003.10.003.
56. Costa, D. A., De Souza, C. J., Costa, P. S., Rodrigues, M. Q., Dos, S. A. F., Lopes,
M. R., & Fietto, L. G., (2014). Physiological characterization of thermotolerant yeast
for cellulosic ethanol production. Applied Microbiology and Biotechnology, 98(8),
3829–3840. https://doi.org/10.1007/s00253-014-5580-3.
57. Choudhary, J., Singh, S., & Nain, L., (2016). Thermotolerant fermenting yeasts for
simultaneous saccharification fermentation of lignocellulosic biomass. Electronic
Journal of Biotechnology, 21, 82–92. https://doi.org/10.1016/j.ejbt.2016.02.007.
58. Chen, C., Wu, L., Cao, Q., Shao, H., Li, X., Zhang, Y., et al., (2018). Genome comparison
of different Zymomonas mobilis strains provides insights on conservation of the
evolution. PLoS One, 13(4), e0195994. https://doi.org/10.1371/journal.pone.0195994.
59. Bergey’s Manual of Systematics of Archaea and Bacteria First published: 17 April 2015.
Online ISBN: 9781118960608| doi: 10.1002/9781118960608.
60. Kluyver, A. J., & Van, N. K., (1936). Prospects for a natural system of classification
of bacteria. Zentralb. Bakteriol. Parasitenkd. Infektionskr. Hyg. Abt., 2(94), 369–403.
61. De Ley, J., & Swings, J., (1976). Phenotypic description, numerical analysis and a
proposal for an improved taxonomy and nomenclature of the genus zymomonas kluyver
and Van Niel 1936. Int. J. Syst. Bacteriol., 26, 146–157.
62. Escalante, A., Giles-Gómez, M., Hernández, G., Córdova-Aguilar, M. S., López-
Munguía, A., Gosset, G., & Bolívar, F., (2008). Analysis of bacterial community during
the fermentation of pulque, a traditional Mexican alcoholic beverage, using a polyphasic
64. Swingsm, J., & De Ley, J., (1977). The biology of zymomonas. Bacteriol. Revs., 41, 1–46.
65. Jeong-Sun, S., Chong, H., Park, H. S., Kyoung-Oh, Y., Jung, C., Kim, J. J., Hong, J. H.,
et al., (2005). The genome sequence of the ethanologenic bacterium Zymomonas mobilis
ZM4. Nat. Biotechnol., 23(1), 63–8. doi: 10.1038/nbt1045.
66. Jiang, X., Dong, D., Bian, L., Zou, D., He, X., & Ao, D., (2016). Rapid Detection of
Candida Albicans by Polymerase Spiral Reaction Assay in Clinical Blood Samples, 7,
1–6. https://doi.org/10.3389/fmicb.2016.00916.
67. Li, L., Wang, X., Jiao, X., & Qin, S., (2017). Differences between flocculating yeast
Author Copy
and regular industrial yeast in transcription and metabolite profiling during ethanol
fermentation. Saudi Journal of Biological Sciences, 24(3), 459–465. https://doi.
org/10.1016/j.sjbs.2017.01.013.
68. Swinnen, S., Goovaerts, A., Schaerlaekens, K., Dumortier, F., Verdyck, P., Souvereyns,
K., Van, Z. G., et al., (2015). Auxotrophic mutations reduce tolerance of Saccharomyces
cerevisiae to very high levels of ethanol stress. Eukaryotic Cell, 14(9), 884–897. https://
doi.org/10.1128/EC.00053-15.
69. Randez-Gil, F., Córcoles-Sáez, I., & Prieto, J. A., (2013). Genetic and phenotypic
characteristics of baker’s yeast: Relevance to baking. Annual Review of Food Science
and Technology, 4(1), 191–214. https://doi.org/10.1146/annurev-food-030212-182609.
70. Bergey’s Manual of Systematics of Archaea and Bacteria, Online © 2015 Bergey’s Manual
Trust. sThis article is © 2005 Bergey’s Manual Trust. Published by John Wiley & Sons,
Inc., in association with Bergey’s Manual Trust. doi: 10.1002/9781118960608.gbm00925.
71. Schatschneider, S., Huber, C., Neuweger, H., Watt, T. F., Pühler, A., Eisenreich, W., &
Vorhölter, F. J. (2014). Metabolic flux pattern of glucose utilization by Xanthomonas
campestris pv. campestris: prevalent role of the Entner–Doudoroff pathway and minor
fluxes through the pentose phosphate pathway and glycolysis. Molecular BioSystems,
10(10), 2663–2676. https://doi.org/10.1039/C4MB00198B.
72. Díaz, V. H. G., & Willis, M. J. (2019). Ethanol production using Zymomonas mobilis:
Development of a kinetic model describing glucose and xylose co-fermentation. Biomass
and Bioenergy, 123, 41–50. https://doi.org/10.1016/j.biombioe.2019.02.004.
73. Fuchino, K., Kalnenieks, U., Rutkis, R., Grube, M., & Bruheim, P. (2020). Metabolic
profiling of glucose-fed metabolically active resting Zymomonas mobilis strains.
Metabolites, 10(3), 81. https://doi.org/10.3390/metabo10030081.
74. Seo, J. S., Chong, H., Park, H. S., Yoon, K. O., Jung, C., Kim, J. J., ... & Kang, H.
S. (2005). The genome sequence of the ethanologenic bacterium Zymomonas mobilis
ZM4. Nature Biotechnology, 23(1), 63-68. https://doi.org/10.1038/nbt1045.
75. Kalnenieks, U., Balodite, E., Strähler, S., Strazdina, I., Rex, J., Pentjuss, A., ... &
Bettenbrock, K. (2019). Improvement of acetaldehyde production in Zymomonas
mobilis by engineering of its aerobic metabolism. Frontiers in Microbiology, 10, 2533.
https://doi.org/10.3389/fmicb.2019.02533.
76. Zhang, K., Lu, X., Li, Y., Jiang, X., Liu, L., & Wang, H. (2019). New technologies provide
more metabolic engineering strategies for bioethanol production in Zymomonas mobilis.
Applied Microbiology and Biotechnology, 103(5), 2087–2099. https://doi.org/10.1007/
s00253-019-09620-6.
77. Magyar, I., & Tóth, T. (2011). Comparative evaluation of some oenological properties
in wine strains of Candida stellata, Candida zemplinina, Saccharomyces uvarum and
Saccharomyces cerevisiae. Food Microbiology, 28(1), 94–100. https://doi.org/10.1016/j.
fm.2010.08.011.
Apple Academic Press
78. Yang, S., Fei, Q., Zhang, Y., Contreras, L. M., Utturkar, S. M., Brown, S. D., ... & Zhang, M.
(2016). Zymomonas mobilis as a model system for production of biofuels and biochemicals.
Microbial Biotechnology, 9(6), 699–717. https://doi.org/10.1111/1751-7915.12408.
79. Swinnen, S., Goovaerts, A., Schaerlaekens, K., Dumortier, F., Verdyck, P., Souvereyns,
K., Van Zeebroeck, G., Foulquié-Moreno, M. R., & Thevelein, J. M. (2015). Auxotrophic
Mutations Reduce Tolerance of Saccharomyces cerevisiae to Very High Levels of
Ethanol Stress. Eukaryotic Cell, 14(9), 884–897. https://doi.org/10.1128/EC.00053-15.
80. He, M. X., Wu, B., Shui, Z. X., Hu, Q. C., Wang, W. G., Tan, F. R., Tang, X. Y., Zhu, Q.
L., Pan, K., Li, Q., & Su, X. H. (2012). Transcriptome profiling of Zymomonas mobilis
Author Copy
under ethanol stress. Biotechnology for Biofuels, 5(1), 75. https://doi.org/10.1186/1754-
6834-5-75.
81. Hacking, A. J., Taylor, I. W. F., & Hanas, C. M. (1984). Selection of yeast able to
produce ethanol from glucose at 40 C. Applied Microbiology and Biotechnology, 19(5),
361–363. https://doi.org/10.1007/BF00253786.
Author Copy
1
Department of Environmental Sciences, University of Kerala,
Thiruvananthapuram, Kerala, India
2
Bioremediation Laboratory, Autonomous University of Coahuila, México
ABSTRACT
3.1 INTRODUCTION
Author Copy
promising alternatives for nonrenewable resources and hence, significant
technologies should be explored to meet world energy demand for low carbon
fuels [1]. Cellulosic biofuels can make a significant contribution to rural
economy [2]. Ethanol from lignocellulosic material is a very cost-effective
alternative. Due to the complex structure and recalcitrant nature of LCB,
the production of bioethanol from lignocellulosic material needs expensive
chemical or physical pretreatments along with enzyme treatment that has a
negative impact on its industrial large-scale production [3, 4].
Pretreatment is necessary to convert cellulosic material to fermentable
form, such treatment enable the feedstock readily available for the enzyme
action [5]. Acid or hydrolytic enzymes can be used as pretreatment, in which
acid or enzyme degrade complex cellulose to glucose monomers by disrupting
the hydrolytic linkage [6]. The main problem associated with the produc-
tion of microbial ethanol from lignocellulosic material is the conversion of
cellulosic material to fermentable form [7]. It is important to overcome these
impacts for an economically feasible sustainable production by integrating
all processes into CBP and genetic manipulation of ethanol producers [8].
This review consolidates the characteristic features and the physiology of C.
thermocellum that makes it an attractive and potential candidate for ethanol
production.
thermophiles for second generation ethanol production was put forth [8]. The
main advantage of thermophiles over mesophiles is that the thermophiles can
utilize a wide range of substrate and allow direct-ethanol recovery through
in situ vacuum distillation. Moreover, they have the advantages of tolerating
Apple Academic Press
Author Copy
of substrates and thrive under mesophilic or thermophilic temperature
conditions [12]. The widely explored Clostridium species is C. thermocellum,
which could grow at a temperature range of 50 and 68°C and are able to
degrade complex cellulosic structures with the presence of cellulosome,
a complex hydrolytic enzyme [13] which has demonstrated remarkable
hydrolysis efficiency of cellulose [14]. Moreover, thermophilic cellulolytic
microorganisms like C. thermocellum are regarded as potential candidates
because their growth at high temperatures reduces the risk of contamination
and also increases the substrate solubility [15].
Clostridium thermocellum is an anaerobic, rod-shaped, Gram-positive
thermophilic microbe which is able to produce ethanol directly from cellu-
lose. C. thermocellum is an ideal bacterium for synthesizing various alcohols
including ethanol and isobutanol [16, 17]. Thermophilic cellulolytic micro-
organisms are well suited for this process because thermophilic condition
reduces the risk of contamination and increases the solubility of feedstock
[13]. C. thermocellum is an ideal candidate for CBP which includes a single
step which includes the degradation of lignocellulosic material and their
conversion to useful products [16]. In 1926, Viljoen et al. [18] attempted
to identify the organism capable of degrading cellulose, which lead to the
basic identification of C. thermocellum. Further, McBee [19] character-
ized C. thermocellum and observed that they can grow in between 50 to
60°C in substrates such as cellulose, cellobiose, and hemicellulose. The
various byproducts were CO2, hydrogen gas, acetic acid, succinic acid, and
ethanol. After these studies, Freier et al. [20] revealed that the optimum pH
and temperature for these organisms are 6 to 7 and 55°C respectively. C.
thermocellum can be cultured in both batch and continuous culture, with
growth rates of 0.10/h and 0.16/h, respectively [21]. The attractive feature
of C. thermocellum is the presence of cellulosome, which is an extracel-
lular multienzyme complex containing around 20 enzymes able to reduce
Author Copy
pathway is controlled by a set of cofactors such as NADH, ATP, etc., through
glycolysis and hexose is converted to pyruvate. After that pyruvate is reduced
to acetyl coA by pyruvate ferredoxin oxidoreductase, that leads to the further
reduction of acetyl coA to ethanol by alcohol dehydrogenase (ADH). Final
end products are mixed acids. The formation of these byproducts is the main
obstacle for high ethanol yield. Understanding the electron transfer mechanism
is very important for the metabolic engineering of desired organism [27].
Author Copy
FIGURE 3.1 Simplified scheme of glucose degradation to various end products by strictly
anaerobic bacteria. [Enzyme abbreviations: ALDH: acetaldehyde dehydrogenase; ADH:
alcohol dehydrogenase; AK: acetate kinase; FNOR: ferredoxin oxidoreductase; H2-ase:
hydrogenase; LDH: lactate dehydrogenase; PFOR: pyruvate ferredoxin oxidoreductase; and
PTA: phosphotransacetylase].
Author Copy
missing components and requires further investigations.
The ideal microorganisms for CBP should have several essential char-
acteristics including simultaneous utilization and conversion of multiple
sugars like cellobiose, glucose, and xylose and also the ability to tolerate
toxic by-products. Most of the studies involving CBP is based on ethanol
fermentation using model substrates like cellulose and xylan and studies
pertaining to real substrate-based ethanol fermentation is limited. Hence,
enhanced ethanol production from native LCB like rice straw biomass is
Author Copy
the goal of sustainable CBP. Co-culture of different cellulolytic and sugar-
fermenting thermophilic anaerobic bacteria has been widely studied to
achieve improved ethanol production using CBP. Singh et al. [40] demon-
strated the direct fermentation ability of a thermophilic anaerobic cellulolytic
bacteria, C. thermocellum DSM 1313 isolated from Himalayan hot spring,
to convert various cellulosic and hemicellulosic substrates into ethanol
using a CBP based approach without addition of any exogenous enzymes.
This strain showed good ethanol production on pretreated rice straw. Use
of CBP in butanol production from LCB is a good option due to the high
energy density of butanol. A co-culture of C. thermocellum and C. saccha-
roperbutylacetonicum has been reported to produce a significant amount of
butanol using crystalline cellulose as substrate [41], and further research is
also focused on developing new pathways of butanol production by genetic
engineering of C. thermocellum.
3.6 CONCLUSION
KEYWORDS
• C. thermocellum
Apple Academic Press
• consolidated bioprocessing
• ethanol production
• lactate dehydrogenase
• phosphotransacetylase
• strain improvement
Author Copy
REFERENCES
1. Lynd, L. R., Liang, X., Biddy, M. J., Allee, A., Cai, H., Foust, T., Himmel, M. E., Lser, M. S.,
et al., (2017). Cellulosic ethanol: status and innovation. Current Opinion in Biotechnology,
45, 202–211.
2. Jordan, N., Boody, G., Broussard, W., Glover, J. D., Keeney, D., McCown, B. H., McIsaac,
G., et al., (2007). Sustainable Development of the Agricultural Bio-Economy, 316(5831),
1570. Science-New York then Washington.
3. Sanchez, O. J., & Cardona, C. A., (2008). Trends in biotechnological production of fuel
ethanol from different feedstocks. Bioresource Technology, 99(13), 5270–5295.
4. Taylor, M. P., Eley, K. L., Martin, S., Tuffin, M. I., Burton, S. G., & Cowan, D. A.,
(2009). Thermophilic ethanologenesis: Future prospects for second-generation bioethanol
production. Trends in Biotechnology, 27(7), 398–405.
5. Gupta, R. B., & Demirbas, A., (2010). Gasoline, Diesel, and Ethanol Biofuels from
Grasses and Plants. Cambridge University Press.
6. Orozco, A., Ahmad, M., Rooney, D., & Walker, G., (2007). Dilute acid hydrolysis of
cellulose and cellulosic bio-waste using a microwave reactor system. Process Safety and
Environmental Protection, 85(5), 446–449.
7. Viikari, L., Vehmaanperä, J., & Koivula, A., (2012). Lignocellulosic ethanol: From science
to industry. Biomass and Bioenergy, 46, 13–24.
8. Scully, S. M., & Orlygsson, J., (2015). Recent advances in second generation ethanol
production by thermophilic bacteria. Energies, 8(1), 1–30.
9. Sveinsdottir, M., Baldursson, S. R. B., & Örlygsson, J., (2009). Ethanol production from
monosugars and lignocellulosic biomass by thermophilic bacteria isolated from Icelandic
hot springs. Icel. Agric. Sci., 22, 45–48.
10. Fong, J. C., Svenson, C. J., Nakasugi, K., Leong, C. T., Bowman, J. P., Chen, B.,
Glenn, D. R., et al., (2006). Isolation and characterization of two novel ethanol-tolerant
facultative-anaerobic thermophilic bacteria strains from waste compost. Extremophiles,
10(5), 363–372.
11. Crespo, C., Pozzo, T., Karlsson, E. N., Alvarez, M. T., & Mattiasson, B., (2012). Caloramator
boliviensis sp. nov., a thermophilic, ethanol-producing bacterium isolated from a hot spring.
International Journal of Systematic and Evolutionary Microbiology, 62(7), 1679–1686.
12. Wiegel, J., Tanner, R., & Rainey, F. A., (2006). An introduction to the family Clostridiaceae.
The Prokaryotes, 4, 654–678.
13. Demain, A. L., Newcomb, M., & Wu, J. D., (2005). Cellulase, clostridia, and ethanol.
Microbiology and Molecular Biology Reviews, 69(1), 124–154.
Apple Academic Press
14. Lu, Y., Zhang, Y. H. P., & Lynd, L. R., (2006). Enzyme - microbe synergy during cellulose
hydrolysis by Clostridium thermocellum. Proc. Natl. Acad. Sci. U.S.A. 103, 16165–16169.
doi: 10.1073/pnas.0605381103.
15. Blumer-Schuette, S. E., Brown, S. D., Sander, K. B., Bayer, E. A., Kataeva, I., Zurawski,
J. V., et al., (2013). Thermophilic lignocellulose deconstruction. FEMS Microbiol. Rev.,
38, 393–448. doi: 10.1111/1574-6976.12044.
16. Akinosho, H., Yee, K., Close, D., & Ragauskas, A., (2014). The emergence of Clostridium
thermocellum as a high utility candidate for consolidated bioprocessing applications.
Frontiers in Chemistry, 2, 66.
Author Copy
17. Holwerda, E. K., Thorne, P. G., Olson, D. G., Amador-Noguez, D., Engle, N. L.,
Tschaplinski, T. J., Van, D. J. P., & Lynd, L. R., (2014). The exometabolome of Clostridium
thermocellum reveals overflow metabolism at high cellulose loading. Biotechnology for
Biofuels, 7(1), 155.
18. Viljoen, J. A., Fred, E. B., & Peterson, W. H., (1926). The fermentation of cellulose by
thermophilic bacteria. The Journal of Agricultural Science, 16(1), 1–17.
19. McBee, R. H., (1954). The characteristics of Clostridium thermocellum. Journal of
Bacteriology, 67(4), 505.
20. Freier, D., Mothershed, C. P., & Wiegel, J., (1988). Characterization of Clostridium
thermocellum JW20. Applied and Environmental Microbiology, 54(1), 204–211.
21. Lynd, L. R., Grethlein, H. E., & Wolkin, R. H., (1989). Fermentation of cellulosic
substrates in batch and continuous culture by Clostridium thermocellum. Applied and
Environmental Microbiology, 55(12), 3131–3139.
22. Bayer, E. A., Henrissat, B., & Lamed, R., (2009). The cellulosome: A natural bacterial
strategy to combat biomass recalcitrance. Biomass Recalcitrance: Deconstructing the
Plant Cell Wall for Bioenergy, 407–435.
23. Bomble, Y. J., Lin, C. Y., Amore, A., Wei, H., Holwerda, E. K., Ciesielski, P. N., Donohoe,
B. S., et al., (2017). Lignocellulose deconstruction in the biosphere. Current Opinion in
Chemical Biology, 41, 61–70.
24. Himmel, M. E., Xu, Q., Luo, Y., Ding, S. Y., Lamed, R., & Bayer, E. A., (2010). Microbial
enzyme systems for biomass conversion: Emerging paradigms. Biofuels, 1(2), 323–341.
25. Pawar, P. M. A., Koutaniemi, S., Tenkanen, M., & Mellerowicz, E. J., (2013). Acetylation
of woody lignocellulose: Significance and regulation. Frontiers in Plant Science, 4, 118.
26. Rydzak, T., McQueen, P. D., Krokhin, O. V., Spicer, V., Ezzati, P., Dwivedi, R. C.,
Shamshurin, D., et al., (2012). Proteomic analysis of Clostridium thermocellum core
metabolism: Relative protein expression profiles and growth phase-dependent changes
in protein expression. BMC Microbiology, 12(1), 214.
27. Lo, J., Olson, D. G., Murphy, S. J. L., Tian, L., Hon, S., Lanahan, A., Guss, A. M., &
Lynd, L. R., (2017). Engineering electron metabolism to increase ethanol production in
Clostridium thermocellum. Metabolic Engineering, 39, 71–79.
28. Shaw, A. J., Podkaminer, K. K., Desai, S. G., Bardsley, J. S., Rogers, S. R., Thorne,
P. G., & Lynd, L. R., (2008). Metabolic engineering of a thermophilic bacterium to
produce ethanol at high yield. Proceedings of the National Academy of Sciences,
105(37), 13769–13774.
29. Herrero, A. A., & Gomez, R. F., (1980). Development of ethanol tolerance in Clostridium
thermocellum: Effect of growth temperature. Applied and Environmental Microbiology,
40(3), 571–577.
30. Argyros, D. A., Tripathi, S. A., Barrett, T. F., Rogers, S. R., Feinberg, L. F., Olson, D.
Apple Academic Press
G., Foden, J. M., et al., (2011). High ethanol titers from cellulose by using metabolically
engineered thermophilic, anaerobic microbes. Applied and Environmental Microbiology,
77(23), 8288–8294.
31. Biswas, R., Zheng, T., Olson, D. G., Lynd, L. R., & Guss, A. M., (2015). Elimination of
hydrogenase active site assembly blocks H2 production and increases ethanol yield in
Clostridium thermocellum. Biotechnology for Biofuels, 8(1), 1–8.
32. Rydzak, T., Lynd, L. R., & Guss, A. M., (2015). Elimination of formate production in
Clostridium thermocellum. Journal of Industrial Microbiology & Biotechnology, 42(9),
1263–1272.
Author Copy
33. Papanek, B., Biswas, R., Rydzak, T., & Guss, A. M., (2015). Elimination of metabolic
pathways to all traditional fermentation products increases ethanol yields in Clostridium
thermocellum. Metabolic Engineering, 32, 49–54.
34. Gaida, S. M., Liedtke, A., Jentges, A. H. W., Engels, B., & Jennewein, S., (2016). Metabolic
engineering of Clostridium cellulolyticum for the production of n-butanol from crystalline
cellulose. Microbial. Cell Factories, 15(1), 1–11.
35. Lin, P. P., Mi, L., Morioka, A. H., Yoshino, K. M., Konishi, S., Xu, S. C., Papanek, B. A.,
et al., (2015). Consolidated bioprocessing of cellulose to isobutanol using Clostridium
thermocellum. Metabolic Engineering, 31, 44–52.
36. Kannuchamy, S., Mukund, N., & Saleena, L. M., (2016). Genetic engineering of Clostridium
thermocellum DSM1313 for enhanced ethanol production. BMC Biotechnology, 16(Suppl
1), 34.
37. Hon, S., Olson, D. G., Holwerda, E. K., Lanahan, A. A., Murphy, S. J. L., Maloney, M. I.,
Zheng, T., et al., (2017). The ethanol pathway from Thermoanaerobacterium saccharolyticum
improves ethanol production in C. thermocellum. Metabolic Engineering, 42, 175–184.
38. Xu, Q., Singh, A., & Himmel, M. E., (2009). Perspectives and new directions for the
production of bioethanol using consolidated bioprocessing of lignocellulose. Current
Opinion in Biotechnology, 20(3), 364–371.
39. Dash, S., Olson, D. G., Chan, S. H. J., Amador-Noguez, D., Lynd, L. R., & Maranas,
C. D., (2019). Thermodynamic analysis of the pathway for ethanol production from
cellobiose in Clostridium thermocellum. Metabolic Engineering, 55, 161–169.
40. Singh, N., Mathur, A. S., Tuli, D. K., Gupta, R. P., Barrow, C. J., & Munish, P., (2017).
Cellulosic ethanol production via consolidated bioprocessing by a novel thermophilic
anaerobic bacterium isolated from a Himalayan hot spring. Biotechnol. Biofuels, 10, 73.
41. Nakayama, S., Kiyoshi, K., Kadokura, T., & Nakazato, A., (2011). Butanol production
from crystalline cellulose by cocultured Clostridium thermocellum and Clostridium
saccharoperbutylacetonicum N1-4. Appl. Environ. Microbiol., 77, 6470–6475. doi:
10.1128/AEM.00706-11.
42. Tomás, A. F. (2013). Optimization of Bioethanol Production from Carbohydrate Rich
Wastes by Extreme Thermophilic Microorganisms (Doctoral dissertation, PhD Thesis,
Technical University of Denmark, Copenhagen, Denmark).
43. Dien, B. S., Cotta, M. A., & Jeffries, T. W. (2003). Bacteria engineered for fuel ethanol
production: current status. Applied microbiology and biotechnology, 63(3), 258–266.
Author Copy
DANIA SANDOVAL-NUÑEZ,1 TERESA ROMERO-GUTIÉRREZ,2
MELCHOR ARELLANO-PLAZA,1 ANNE GSCHAEDLER,1 and
LORENA AMAYA-DELGADO1
1
Industrial Biotechnology Unit, Center for Research and Assistance in
Technology and Design of the State of Jalisco A.C., Camino Arenero–1227,
El Bajio del Arenal, C.P.–45019, Zapopan, Jalisco, Mexico
2
Computer Sciences Department, Exact Sciences and Engineering
University Centre, Universidad de Guadalajara, Blvd. Gral. Marcelino
Garcia Barragan–1421, Olimpica. Guadalajara, Mexico
ABSTRACT
Interest in ethanol production has been increasing since the 1980s because it is
considered an alternative source of clean energy. The search for new and cheaper
raw materials drove the development of new processes to obtain ethanol, such
as the fermentation of lignocellulosic hydrolysates. In lignocellulosic ethanol
production, microbial cells are under a stress-inducing environment, and
their physiological behavior is altered significantly, indicating that different
gene regulatory mechanisms are activated from those in non-stress-inducing
conditions. Ethanologenic yeasts and bacteria tune gene expression to regulate
response mechanisms and quickly adapt their global metabolic networks to
grow and produce ethanol under unfavorable circumstances. Understanding
the regulatory mechanism of gene expression during ethanol production has
been a main topic in molecular and cellular biology with the goal of developing
more efficient and resistant ethanol-producing microorganisms.
thermocellum.
4.1 INTRODUCTION
Author Copy
yeast Saccharomyces cerevisiae, which is capable of fermenting C6 sugars
only [1]. However, the degradation of other materials such as lignocellulosic
biomass (LCB) also produces C5 sugars; this problem has been mitigated by
using recombinant strains of S. cerevisiae and bacteria such as Z. mobilis and
some thermophiles such as Clostridium thermocellum. Interestingly, micro-
organisms such as C. thermocellum can not only efficiently degrade cellulose
and hemicellulose but also ferment hexose sugars to ethanol. Regardless
of the microorganism used for ethanol production, fermentation efficiency
depends on cellular metabolism, which is directly related to gene regulation.
Genetic regulation is essential for microorganisms in several biological
processes, including the generation of biomolecules such as ethanol. During
ethanol production, gene regulation plays a critical role in the cellular
capacity to adapt quickly to the physicochemical conditions of the process,
the ability to utilize carbon sources, and the ability to use alternative
metabolic pathways to overcome nutrient-limiting conditions or respond to
stressors [2]. Genetic regulation can be exerted on different levels in the cell;
the most basic level is transcription. Regulation in eukaryotic cells requires
the coordination of a whole set of genes that are scattered over the genome.
This mechanism is different in bacteria, where genes, which are regulated
in the same way, are often organized into operons, and transcribed from one
regulatory sequence. When ethanol is produced, two types of regulation can
be present: stimulation of the expression of specific genes (positive regula-
tion) or inhibition of their expression (negative regulation) [1, 2]. Several
studies have demonstrated that some restrictions are present in bioethanol
production when wild-type microorganisms are used. These restrictions
involve different biotic (formation of unwanted products) and abiotic effects
(source of carbon, nitrogen, nutrients, temperature, oxygen, and the presence
of stressful molecules) on the microorganisms. These restrictions may lead
to a low conversion rate and low yield in industrial ethanol production. Thus,
Author Copy
for the positive or negative regulation of specific genes that increase ethanol
yields. This chapter presents the state of knowledge in the field of genetic
regulatory mechanisms in the most important microorganisms used to
produce ethanol. It is based on experimental evidence of how microorganisms
combine different regulatory mechanisms to coordinate multiple metabolic
pathways during ethanol production.
tory mechanisms carried out in yeasts during the fermentation process. This
knowledge allows control of the fermentation processes to achieve higher
yields and productivities. Some of the primary yeast taxa most studied in
ethanol production processes are S. cerevisiae and K. marxianus.
One of the main challenges faced in the fermentation process for the
production of 2G ethanol is the presence of various toxic compounds that
affect yeast physiology and metabolism, inhibiting or reducing ethanol yields.
S. cerevisiae has been studied in several fermentation processes where stress
Apple Academic Press
factors such as high temperatures, the absence of nutrients, and the presence
of cell growth inhibitors such as organic acids, furans, phenols, and aldehydes
intervene [6–9]. The presence of inhibitory compounds causes changes in
yeast gene regulation, mainly during the transcription process, where the
overexpression or inactivation of genes is involved in the consumption of the
carbon source and therefore in ethanol production. However, the glycolysis
pathway is not the only pathway affected by the presence of inhibitors, since
Author Copy
for the carbon source to be consumed and subsequently converted to ethanol,
the inhibitors present in the medium must be metabolized by yeast [10, 11].
Once the inhibitors are metabolized, the yeast is able to consume the sugars
quickly.
Nevertheless, inhibitors activate a series of defense mechanisms and
homeostatic balance in yeasts, achieving an ideal intracellular balance for
ethanol production [9]. The detoxification process in S. cerevisiae employs
a series of oxidation-reduction reactions to remove the high concentration
of reactive oxygen species (ROS) in the mitochondria. On the other hand,
S. cerevisiae uses a series of metabolic pathways to synthesize amino acids,
transporters, and membrane lipids, produce secondary metabolites, and
overexpress enzymes that facilitate the detoxification process and ethanol
generation (Figure 4.2).
under conditions in which none of the other four ADHs (ADH1 to ADH4)
are functional. In fermentation processes in the presence of furan aldehydes
(furfural and HMF), the overproduction of ATP and NADPH has been
observed since the responses to stress by furan aldehydes generate oxida-
tive stress, which provokes a considerable increase in the net production of
NADPH in strains of S. cerevisiae (> 4-fold). In this condition, the additional
NADPH could be used by ADHs (ADH6 and ADH7) to convert furan alde-
Author Copy
hydes into less toxic compounds, lowering the toxicity of the medium for the
yeast [12].
FIGURE 4.2 Schematic representation of the modulation pathways and genes involved
in ethanol production under stress conditions. [Abbreviations: HXK: hexokinase; GLK:
glucokinase; MDH: malate dehydrogenase; THI: thiamine metabolism regulatory protein; PDR:
transcription factor; YOR: oligomycin resistance ATP-dependent permease; STB: protein STB,
DNA-binding transcription factor; YAP: AP-1-like transcription factor; GZF: DNA-binding
transcription factor; LEU: regulatory protein; PUT: proline utilization trans-activator; WAR: weak
acid resistance protein; SFA: S-(hydroxymethyl)glutathione dehydrogenase; ALD4: potassium-
activated aldehyde dehydrogenase; ALD6: magnesium-activated aldehyde dehydrogenase;
ADH: alcohol dehydrogenase; RHR: glycerol 3-phosphate phosphohydrolase; HOR: glycerol-
1-phosphate phosphohydrolase; CHO: phosphatidylethanolamine N-methyltransferase; ARI:
NADPH-dependent aldehyde reductase; OYE: NADPH dehydrogenase; ALD: aldehyde
dehydrogenase; TFs: transcription factors].
Author Copy
In addition to the study of the expression of various critical genes involved
in ethanol production, transcription factors (TFs) play an essential role in the
production of 2G ethanol, participating in processes of adaptation and yeast
resistance to stressful conditions. TFs such as YAP1, STB5, DAL81, GZF3,
LEU3, PUT3, and WAR1 have been widely studied because they are associ-
ated with various stress response mechanisms. These TFs confer appropriate
biological characteristics on yeast to carry out ethanol production [15]. For
example, YAP1 is associated with the glutathione pathway and is essential in
the intracellular detoxification of ROS. On the other hand, some TFs, such as
LEU3 and DAL81, are involved in carbon metabolism, amino acid biosyn-
thesis, and nitrogen catabolism, allowing yeasts to use the available energy
in the form of ATP and NADPH to carry out detoxification processes [7, 15].
FIGURE 4.3 Multiple sequence alignment analysis of the alcohol dehydrogenases (ADH1 to ADH7) identified in Saccharomyces cerevisiae.
The conserved residues are highlighted in purple. Residues boxed in red correspond to the catalytic zinc-binding motif. Residues boxed in orange
belong to the binding site with NADH/NADPH. The multiple sequence alignment was obtained with the MUSCLE algorithm (https://www.ebi.
Author Copy
Genetic Regulation of Principal Microorganisms 61
Author Copy
ionic stress. On the other hand, SKN7 forms part of the SLN1-YPD1-SNN7
two-component regulatory system, which is related to the expression control
of genes involved in the response to changes in extracellular osmolarity.
As mentioned above, gene regulation occurs most frequently at the
transcriptional level. The regulation of gene expression in S. cerevisiae
under stress-inducing conditions can affect several biological processes and
trigger a series of regulatory events mediated by numerous TFs (Figure 4.4).
Genes related to the processes of detoxification, adaptation, and tolerance in
S. cerevisiae are regulated by key TFs, such as the YAP gene family, PDR
gene family, RPN4, and HSF1 [7, 9, 15].
The primary genes activated during 2G ethanol production in the pres-
ence of inhibitors such as furfural and HMF are described below.
In recent years, various studies have been carried out to find strategies
to understand the molecular mechanisms involved in the adaptation and
tolerance of S. cerevisiae with respect to the main groups of growth and
fermentation inhibitors, such as aldehydes, phenols, ketones, and weak
organic acids. Furfural, HMF, and acetic acid are the three main compounds
that participate in the inhibition of the 2G ethanol production process [17].
For ethanol production, fermentation is carried out at pH values between
4 and 5, causing the dissociation of acetic acid (pKa = 4.76), which diffuses
through the plasma membrane of yeast [18]. Inside the cell, the acid dissoci-
ates and releases a proton, making it necessary to use ATP to reach the cell
equilibrium required for essential biological functions and prevent cytosolic
acidification. The use of ATP causes a decrease in yeast growth during the first
hours of contact; this condition remains until the yeast manages its genetic
machinery to adapt to this adverse condition. The acetic acid concentration
can be higher than 5 g/l in lignocellulosic residue hydrolysates, increasing
the inhibition problem caused by acetic acid. At a high concentration of
Author Copy
most important in the response of yeast to acetic acid. Other important
factors in resistance to acetic acid stress are Rim101p and Msn2p, which are
also displayed under other stress types, such as ethanol, high concentrations
of glucose and oxidative stress [21–23].
Furfural is well known to be transformed into furfuryl alcohol by
nicotinamide adenine dinucleotide-dependent dehydrogenases (NADH)
under anaerobic conditions. Therefore, enzymes that participate in the
reduction of aldehyde groups are essential to reduce the concentration of
inhibitors in lignocellulosic residues. The genes reported to participate in
the detoxification of furfural and HMF are ADH6, ADH7, ALD4, GRE3 [9];
ADH1 [24]; ARI1 [25]; and GRE2 [26]; in addition, the Y62 and Y76 genes
participate only in the detoxification of furfural, but no evidence has been
found that they also participate in HMF. Studies developed to evaluate the
capacity of ADH6 and ADH7 in the reduction of furfural and HMF show that
these enzymes have a high capacity to reduce aldehydes, as much as 100 times
greater than their oxidation capacity [27]. The synergistic activity between the
aldehyde dehydrogenase enzymes Ald5, Pad1 decarboxylase, and the alcohol
acetyltransferases Atf1 and Atf2 has been determined to provide resistance
to phenolic inhibitors [28]; thus, it is likely that they participate actively in
the transformation of furans present in lignocellulosic residues. Ma and Liu
demonstrated that transcriptional genes such as YAP1, PDR1, PDR3, RPN4
and HSF1 actively participate in the adaptation of S. cerevisiae to HMF [29].
In recent years, the existence of short-chain aldehyde reductase enzymes
known as SDR has been evidenced; SDR enzymes participate in the
transformation of aldehydes within yeast cells, including Ykl107wp. This
enzyme can reduce acetaldehyde, glycolaldehyde, furfural, formaldehyde,
HMF, and propionaldehyde, but it was not observed to reduce the six ketones
corresponding to the same compounds [30]. Aldehyde reductase enzymes
are highly diverse. SDRs have particular substrate preferences and are found
in different places within the cell. In addition, considering that aldehydes can
cross mitochondrial, nuclear, and cytoplasmic membranes, the presence of
specific aldehyde reductases in each zone guarantees a decrease in cellular
damage [31, 32].
Apple Academic Press
Author Copy
because, during evolution, strains can undergo the necessary changes to allow
higher tolerance to the inhibitors present in the lignocellulosic hydrolysates
and even increase ethanol production and productivity [33–35].
Author Copy
Despite all the successful results obtained through the years in under-
standing stress response mechanisms in S. cerevisiae, it is evident that a
rational strategy is necessary to improve yeast tolerance of ethanol or inhibi-
tory compounds. Such a strategy should combine omic analyzes (genomics,
transcriptomics, proteomics, metabolomics, and fluxomics) with predictive
computational modeling or simulation to design tolerant yeasts with better
metabolic properties for industrial ethanol production.
TABLE 4.1 Ethanol Production by Kluyveromyces marxianus Using Different Carbon Sources
Raw Material Strain Fermentation Process Ethanol (g/l h) References
Barley straw K. marxianus CECT 10875 SSF 0.40 [44]
Sugar cane juice K. marxianus DMKU 3-1042 Batch 1.30 [45]
Wheat straw slurry K. marxianus CECT 10875 Batch 1.69 [46]
Wheat straw slurry K. marxianus CECT 10875 Batch 0.36 [47]
YPD K. marxianus BUNL‑21 Batch 0.06 [48]
Sweet sorghum juice K. marxianus DBKKUY-103 Batch 1.42 [49]
Wheat straw K. marxianus DBTIOC-35 SSF 0.86 [50]
YPDX K. marxianus CBS712 Batch 1.18 [51]
Kanlow switchgrass K. marxianus IMB4 SSF 0.23 [52]
Wheat straw K. marxianus CECT 10875 SSF 0.50 [53]
Corncob residue K. marxianus NBRC1777 Batch 1.05 [54]
Sugar cane bagasse K. marxianus NRRL Y-50883 (SLP1) Batch 0.29 [55]
Mineral medium K. marxianus NRRL Y-50883 (SLP1) Batch 0.30 [10]
Author Copy
Genetic Regulation of Principal Microorganisms 67
at all stages of the cell cycle. However, several extracellular factors contribute
to the use of regulatory mechanisms associated with stress responses and cell
repair. The participation of some TFs, the overexpression and inactivation
of genes, and the synthesis and degradation of some metabolites have
been demonstrated when K. marxianus is employed in ethanol production
processes under certain fermentation conditions in the presence of different
carbon sources, nutrients, and toxic molecules [10, 42, 56].
Author Copy
Ethanol production by K. marxianus has been evaluated mainly with
the use of agro-industrial residues and in defined culture media. Therefore,
understanding some of the molecular regulatory mechanisms employed by
K. marxianus during ethanol production is of great interest. In addition to
the carbon source used and its subsequent conversion to ethanol, several
biochemical and metabolic biological events disrupt ethanol conversion
pathways and mechanisms. Environmental or chemical effects provoke
biological responses by yeasts and consequently intervene in gene regulation
during the ethanol production process, inhibiting or decreasing the genera-
tion of ethanol [56, 58].
The presence of toxic molecules such as furfural and organic acids plays
an important role in gene regulation, biochemically affecting ethanol produc-
tion due to the overexpression of proteins and genes, affecting metabolic
pathways such as glycolysis, the pentose phosphate pathway (PPP), gluta-
thione, and butanoate metabolism, fatty acid metabolism, and the biosyn-
thesis of many amino acids such as Ala, Arg, Asp, Glu, His, Ile, Leu, Pro,
Trp, Tyr, and Val, in addition to MAPK signaling pathways. Some metabolic
pathways, such as those involving secondary metabolites and the synthesis,
and degradation of amino acids, TCA, and glyoxylate, act as intermediaries
in regulatory processes related to defense mechanisms and homeostasis,
maintaining the ideal intracellular balance for ethanol production [56, 59].
Author Copy
Studies focused on ethanol production from xylose as a carbon source
have also achieved improvement by overexpressing enzymes through HR.
For example, the overexpression of enzymes such as xylose reductase
heterologous (XYL1) and xylitol dehydrogenase (XYL2) have increased the
participation of the phosphate pentose pathway and the flow into ethanol
production [63]. The increase in ethanol yields has been achieved through
the overexpression of native xylokinase (XYL3), L-ribulose-5-phosphate
4-epimerase (RPE1), ribose-5-phosphate isomerase (RKI1), transcetolase
(TKL1), and transaldolase (TAL1) genes as well as pyruvate decarboxylase
(PDC1) and ADH2. The XYL1 and XYL2 heterologous genes were selected
because of their preference for NADP(H) over NAD(H), thus helping to
rectify an imbalance in cofactors during growth on xylose [64].
Several genes and their association with the ethanol production process
have been reported to be involved in posttranslational regulation. In ethanol
production under certain stress conditions (temperature, ethanol, and
chemical agents), protein folding and the expression or repression of genes
coding for carbon source carriers and defense mechanisms have been found
to affect oxidation-reduction processes and intracellular transport, in addi-
tion to genes involved in the inhibition of routes and signaling pathways
associated with regulatory processes in ethanol production (Figure 4.5).
Assays performed in K. marxianus demonstrated the biological importance
of genes that code for intermediary enzymes in the ethanol production
process, mainly enzymes involved in carbon metabolism for both xylose and
glucose, enzymes involved in transport processes, mitochondrial enzymes,
Author Copy
FIGURE 4.5 Interactome of genes involved in ethanol production in Kluyveromyces
marxianus. [Abbreviations: DAK: dihydroxyacetone kinase; GUT: glycerol-3 phosphate
dehydrogenase; HXK: hexokinase; TSA: peroxiredoxin; SDH: succinate dehydrogenase;
ICL: isocitrate lyase; FBA: fructose-biphosphate aldolase; TDH: glyceraldehyde-3-phosphate
dehydrogenase; FBP: fructose-1,6-biphosphatase; KGD: 2-oxoglutarate dehydrogenase
complex; PGI: glucose-6-phosphate isomerase; TPI: triosephosphate isomerase; ADH: alcohol
dehydrogenase; GPM: phosphoglycerate mutase; CIT: citrate synthase; PFK: ATP-dependent
6-phosphofructokinase; PGK: phosphoglycerate kinase; MDH: malate dehydrogenase;
CDC: pyruvate kinase; ENO: enolase; LAT: dihydrolipoyllysine-resiude acetyltransferase
component of pyruvate dehydrogenase complex; ACO: aconitate hydratase; IDH: isocitrate
dehydrogenase; MLS: malate synthase; PCK: phosphoenolpyruvate carboxykinase;
ALD: aldehyde dehydrogenase; MAE: NAD-dependent malic enzyme; GDH: glutamate
dehydrogenase; GND: 6-phosphogluconate dehydrogenase]. The interactome was based on
S. cerevisiae using STRING: functional protein association networks (https://string-db.org/)].
Author Copy
associated with cellular processes carried out in the stationary phase under
anaerobic conditions. The metabolic function of KmADH7 is associated
with the oxidation of hemiacetal as an alternative route for the synthesis of
ethyl acetate, and its expression increases in the stationary phase [65].
Multiple sequence alignment analysis of the ADHs of K. marxianus
shows the similarities and differences between the main four KmADHs. The
similarity percentages of KmADH2 to KmADH4 with respect to KmADH1
are 82, 79, and 78% (Figure 4.6). Similar to that in S. cerevisiae, the amino
acid sequence reveals the metal binding site (zinc catalytic site) and the
NAD-binding site (according to NADH or NADPH dependence). Expres-
sion analysis demonstrated that the expression of KmADHs depends on the
growth phase and carbon source [65].
FIGURE 4.6 Multiple sequence alignment analysis of the alcohol dehydrogenases (KmADH1 to KmADH4) identified in Kluyveromyces
marxianus. The conserved residues are highlighted in purple. Residues boxed in red correspond to the catalytic zinc-binding motif. Residues
boxed in orange belong to the binding site with NADH/NADPH. The multiple sequence alignment was obtained with the MUSCLE algorithm
(https://www.ebi.ac.uk/Tools/msa/muscle/), using the ADHs sequences of Kluyveromyces marxianus (strain ATCC 12424 for KmADH1-2, and
strain DMKU 3-1042 for KmADH3-4). Alignments were edited with Jalview software (https://www.jalview.org/).
Author Copy
72 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
Author Copy
is another metabolite that is recognized as a protein and plasma membrane
protector [67]. Flores-Cosío et al. observed an increase in the concentration
of phosphatidylethanolamine and a decrease in phosphatidylcholine when
K. marxianus was stressed by furan aldehydes, illuminating the effect of
these inhibitors on the plasma membrane and the capacity of K. marxianus to
reorganize its membrane [11]. Despite the potential of metabolomic analyzes
as tools to study the effect of inhibitory compounds on cellular metabolism
and physiology, this omic technique has been rarely used in K. marxianus, so
there are few studies that help to understand the stress response in this yeast
through its metabolome.
Recent studies have demonstrated that K. marxianus is a promising yeast
for ethanol production and other interesting metabolites, such as enzymes,
flavor, and fragrance compounds, and xylitol. K. marxianus has advantages
over other yeasts because it can assimilate several carbon sources (glucose,
xylose, arabinose, sucrose, inulin, fructans, among others). In addition, K.
marxianus is a thermotolerant yeast with a high growth rate and high resistance
to inhibitory compounds, and it can quickly adapt its metabolic machinery
to survive toxic environments. The main obstacle to its development at an
industrial level has been the limited knowledge of its genetics and physiology,
but this is rapidly changing thanks to the new omic technologies, making K.
marxianus a promising yeast for ethanol production.
adjust the metabolic pathways depending on the carbon source, pH, tempera-
ture, toxic compounds, and other nutrients available. Different mechanisms
regulate genes in bacteria and yeasts; bacterial genes are organized into
operons or clusters of coregulated genes. In addition to being physically
Apple Academic Press
close in the genome, these genes are regulated to turn on or off together. This
characteristic of grouping related genes under a common control mechanism
allows bacteria to adapt rapidly to changes in the environment.
Transcription has three steps in bacteria: first, RNA polymerase binds to a
promoter site on DNA to form a closed complex; then, RNA polymerase starts
transcription by opening the DNA duplex to form a transcription bubble. In
the second stage, termed elongation, the transcription bubble moves along
Author Copy
DNA, and the RNA chain is extended by adding nucleotides in the 5’ to 3’
direction. Finally, transcription stops, and the DNA duplex reforms when
RNA polymerase dissociates at a terminator site. In bacteria, each gene or
operon is flanked by a promoter and a terminator. The promoter is a specific
nucleotide sequence site where the RNA polymerase binds to DNA and starts
making RNA (mRNA). The terminator is a similar instruction in the DNA
where the RNA polymerase stops transcribing mRNA and dissociates from
the DNA. This mechanism is the purest form of gene expression regulation
in bacteria. Essential components of transcription are sigma factors (σ),
which are subunits of all bacterial RNA polymerases. They are responsible
for determining the specificity of promoter DNA binding and efficiently
control transcription initiation. In conclusion, the first step in bacterial gene
expression and the step most often controlled is transcription. Regulatory
factors usually determine whether a specific gene is transcribed by RNA
polymerase or not under specific environmental conditions.
to low ATP production compared with the EMP pathway for glycolysis, but
it is also present in anaerobic bacteria such as E. coli and Z. mobilis, and it
produces one ATP molecule per glucose consumed [69]. In this pathway,
two key enzymes participate, PDC and two ADHs, which convert pyruvate
Apple Academic Press
Author Copy
expression systems, the silencing of specific genes, and mutagenesis, among
others [73]. One example is a strain of Klebsiella oxytoca that was modified by
metabolic engineering by inserting the PDC gene from Z. mobilis to increase
its capacity for pyruvate decarboxylation, a key enzyme in the homoethanol
pathway of Z. mobilis [74]. In recent years, next-generation sequencing
(NGS) technologies have allowed the characterization of complete genomes
from different microorganisms of industrial interest, including Z. mobilis.
Knowledge of the genetic structure of these bacteria has allowed us to deeply
study the production of metabolites with biotechnological applications and
the adaptation mechanisms to environmental factors. The first genome of Z.
mobilis was sequenced in 2005, with a circular chromosome with a length
of 2,056,416 bp and 5 plasmids, and it was further annotated in 2009 [75,
76]. Subsequently, other genomes were also sequenced with an approximate
length of 2.01 to 2.22 Mb including 2 to 6 plasmids [73].
(Table 4.2) [77, 78]. The gene expression dynamics in Z. mobilis ZM4 were
also evaluated under aerobic and anaerobic conditions. In the absence of
oxygen, several overexpressed genes participating in the ED pathway that
optimize glucose metabolism were identified. Additionally, overexpressed
Apple Academic Press
Author Copy
TABLE 4.2 Genes of Zymomonas mobilis Z4 Differentially Expressed Under Different
Stress Conditions (Temperature and Aeration)#
Heat Stress
Gene ID Cellular Process References
sod, cat, ZMO1573, ZZ6-0186, ahpc Oxidative stress [77]*
dnaKJ, hsp20, clpB, clpA, clpS HSP genes [78]*
[75]*
Aeration conditions
Anaerobic conditions
glk, zwf, pgl, pgk, eno, pdc, adhB ED pathway [79]
rars1 Ribosome-mediated polypeptide
synthesis
leuC, trpB, argC, ilvI, ilvC, thrC, thiC, Amino acid and co-factor
and ribC biosynthetic genes
Aerobic conditions
ZMO0084, ZMO0641, ZMO0651 Chemotaxis [79]
ZMO1022, ZMO1460, NT01ZM1467 Metabolism of sulfur compounds
ZMO1121, ZMO1216, ZMO1387, Transcriptional regulators
ZMO1063
nadE Nicotinamide adenine dinucleotide
de novo biosynthesis
atpA, atpB ATPase alpha/beta chains family
tdsD, nifF Flavoprotein transcripts
ZMO1097, ZMO1830, ZMO1732, Stress response
ZMO0279, ZMO1118, ZMO0749
#
Genes highlighted in green are up regulated and those selected in red are down regulated. The
experiments highlighted with (*) were performed by RT-qPCR.
Author Copy
biofuels and have been reported as toxic agents that can affect cell growth and
fermentation [81]. Yi et al. evaluated the genomic response of Z. mobilis ZM4
in the presence of the inhibitors 4-hydroxybenzaldehyde, syringaldehyde,
and vanillin, identifying overexpressed genes from the respiratory chain and
transporter genes (Table 4.3) that help reduce inhibitors to their corresponding
phenolic alcohols and maintain ethanol production [82].
tolerance were identified as both up- and downregulated and are related to cell
motility and cell wall membrane biogenesis: Several downregulated genes
are implicated in protein synthesis and DNA replication, recombination,
and repair, and upregulated genes are related to transcriptional regulation
in response to DNA damage, which has been extensively studied in other
bacteria, e.g., Escherichia coli and fermentative yeasts [84, 85].
Author Copy
4.3.1.3 GENETIC REGULATION OF ZYMOMONAS MOBILIS UNDER
ETHANOL STRESS
and the components themselves. With this approach, we can develop predic-
tive mathematical models of the biological processes of Z. mobilis.
TABLE 4.4 Genes of Zymomonas mobilis Z4 Differentially Expressed Under Ethanol Stress*
Ethanol
Gene ID Cellular Process References
Author Copy
gnl ED pathway [86]
gntk Pyruvate biosynthesis
ldhA Ethanol formation
flhAB, fliDEFGHIKLMNPQRS, ZMO0613, Cell motility and
ZMO0614, ZMO0604, ZMO605, ZMO607, cell wall membrane
ZMO608, ZMO609, ZMO610, ZMO611, ZMO612, biogenesis
ZMO619, ZMO624, ZMO632, ZMO634, ZMO635,
ZMO642, ZMO643, ZMO648, ZMO649, ZMO651,
ZMO652
ZMO1311
ZMO0022, ZMO1571, ZMO1572, ZMO1032, Respiratory chain
ZMO1255, ZMO1256, ZMO1189, ZMO1669,
ZMO0678, ZMO1812, ZMO1813, ZMO1814, rnfAB
ZMO1844, ZMO0957, ZMO0958, ZMO1252,
ZMO1253, ZMO1254, ZMO1479, ZMO1480,
ZMO1113, ZMO1885, ZMO1809, ZMO1810,
ZMO1811, ZMO0569
ZMO1404, ZMO1623, ZMO0274, ZMO0626 Stress response
ZMO1356, ZMO1426, ZMO1484, ZMO1417, DNA replication,
ZMO1648, ZMO1193, ZMO1401, ZMO086, recombination, and
ZMO0598, ZMO1907, ZMO1187, ZMO1054, repair
ZMO1584, ZMO812
ZMO0054, ZMO2033, ZMO1697, ZMO0281, Transcriptional
ZMO1547, ZMO0774, ZMO0190 regulation
ZMO1107, ZMO0347
ZMO1180, ZMO2018, ZMO1395, ZMO1804, Transport systems
ZMO1025, ZMO1855, ZMO1522, ZMO1425,
ZMO1647, ZMO1262, ZMO0546
ZMO1649, ZMO1757, ZMO0899
*
Genes highlighted in green are up regulated and those selected in red are down regulated.
extracellular free and multienzyme complex termed the cellulosome and its
accessory enzymes. A model of the multienzymatic systems is represented
in Figure 4.8; some proteins were renamed as follows: CipA (ScaA), OlpB
(ScaB), Orf2p (ScaC), OlpA (ScaD), SdbA (ScaF), and OlpC (ScaG) [87].
This enzymatic complex gives C. thermocellum the ability to solubilize the
cellulose contained in LCB and rapidly ferment it to produce ethanol. The
metabolic ability of C. thermocellum to produce ethanol directly from LCB
Author Copy
makes it the main candidate microorganism for bioethanol production via
consolidated bioprocessing (CBP). Despite the biotechnological potential
of C. thermocellum, the industrial application of this bacterium is relegated
because of its disadvantages compared to other microorganisms, such as
mixed acid fermentation, low ethanol productivity and titer, low ethanol
tolerance, and low hemicellulose utilization [88–90]. To solve these biotech-
nological limitations, several strategies have been carried out to improve
ethanol production by C. thermocellum, from technological strategies such
as cocultivation with other bacteria to genomic strategies such as metabolic
engineering. In parallel, several genetic and transcriptomic studies have been
performed to understand the gene regulation involved in biomass degradation,
as well as the ethanol tolerance mechanism used by C. thermocellum [91].
orf4-manB-celT was proposed by Choi et al.; in this model, the protein GlyR3
coregulates the expression of the celC and manB genes [96]. In the extended
model, the expression of the cellulosomal family 26 glycoside hydrolase ManB
is repressed by a high concentration of the protein GlyR3 in the presence of
laminaribiose. In contrast, at a basal concentration of GlyR3 (in the absence
of laminaribiose), the manB gene is expressed (Figure 4.9). The mechanism of
regulation of the celT gene that encodes a cellulosomal family 9 endoglucanase
Author Copy
is unknown. In the same way, the mechanism regulating the orf4 gene and the
function of the protein it encodes are unknown.
Author Copy
Genetic Regulation of Principal Microorganisms 83
binding module known for a strong affinity for cellulose (Figure 4.8).
Ortiz de Ora et al. demonstrated that 5 σI factors (identified as σI1, σI2,
σI3, σI4 and σI6) regulate the expression of 17 genes encoding different cellu-
losomal proteins [98]. They could relate the σI1–σI6 factors to the genetic
regulation of two of the most important components of the cellulosome,
the primary scaffoldin (CipA) and the most abundant enzyme (Cel48S). In
the same way, the regulons of the σI alternative factors were identified by
Author Copy
bioinformatic analysis in conjunction with classical microbiology genetic
tools and the application of the heterologous B. subtilis host system [98].
All σI factor genes (sigl1-sigl6 and sig24C) are upregulated in the presence
of extracellular polysaccharides. In this context, the proposed mechanism to
activate alternate σI implied an extracellular carbohydrate-active module and
an intracellular anti-σ peptide domain [97] (Table 4.5).
Studies performed to understand the mechanisms of genetic regulation
of cellulosome production in C. thermocellum demonstrated that growth
conditions regulate the expression of cellulosomal components and that this
regulation is a response to the polysaccharides present in the culture media.
Nataf et al. identified five sugar ABC transport systems: four are specific
for β-1,4-linked glucose oligomers (cellodextrins), and one is specific for
β-1,3-linked glucose dimer (laminaribiose) [99]. The sugar transporters and
their substrate specificities demonstrated that C. thermocellum prefers to
assimilate cellodextrins rather than cellobiose or glucose. Genome analysis
also suggests that the bacterium lacks any other sugar ABC transporters, in
agreement that this strain can grow only on β-glucans. Consequently, the
sugars present in the culture media determine the composition of the cellulo-
some, which that subsequently influences the overall ability of the bacteria
to degrade lignocellulosic substrates and produce ethanol.
Author Copy
Genetic Regulation of Principal Microorganisms 85
Author Copy
present [104]. The final steps of the EMP pathway involve the successive
reduction of acetyl-CoA and acetaldehyde with electrons provided by NADH
(i.e., the ALDH and ADH reactions); these two reactions are both catalyzed
by the bifunctional alcohol dehydrogenase (ADHE) enzyme. This enzyme
is related to the ethanol tolerance of C. thermocellum, as mutations in the
bifunctional ADHE result in increased ethanol tolerance [101]. Previous
work demonstrated that the expression of ADH genes, pyruvate ferredoxin
oxidoreductase genes (Cthe_2390, Cthe_2391, Cthe_2392 and Cthe_0340)
and other genes related to ethanol production were affected by the carbon
source (cellulose or cellobiose) and the growth rate [105, 106].
Recently, global transcriptomic analysis of C. thermocellum ATCC 27405
growing in dilute acid-pretreated Populus and switchgrass showed overexpres-
sion in genes related to nitrogen uptake and metabolism at the end of fermenta-
tion when populus was used; this overexpression coincided with increases in
ethanol concentration [107]. Similar results were obtained when C. thermo-
cellum was grown on crystalline cellulose [108]. These results demonstrated
that the expression of genes related to ethanol production is influenced by the
carbon source and the fermentation time, as well as the growth rate.
Although several studies have been carried out to understand the mecha-
nism of genetic regulation in one of the most promising bacteria to produce
ethanol directly to LCB, C. thermocellum, substantial further effort is
required to fully elucidate all the genetic regulatory mechanisms that control
the expression of proteins involved in biomass degradation and ethanol
production. The application of new omics technologies (genomics, tran-
scriptomics, proteomics, and metabolomics) allowed us to reach a complete
understanding of metabolism and its connection with the genetic regulation
of C. thermocellum. An understanding of the genetic regulatory mechanisms
in C. thermocellum can help improve this industrially relevant strain and
promote the use of cellulosome-producing bacteria in the CBP of biomass.
Author Copy
temperatures [109].
Extreme thermophilic anaerobic bacteria have been isolated from
different environments, including geothermal areas, volcanic mud, and
canned products [110–112]. These bacteria are facultatively anaerobic
and tolerate extreme pH and salt concentrations during fermentation with
minimal nutrient supplementation [113]. Despite the numerous advantages
of employing thermophilic bacteria in fermentation for bioethanol produc-
tion, it is known that yeasts and other bacteria such as Z. mobilis can tolerate
a higher ethanol concentration than extremophiles due to the fatty acid
composition of the cell membrane [114].
Several studies have been performed with hemicellulolytic thermophiles
of the genera Thermoanaerobacter and Thermoanaerobacterium to increase
the ethanol yield and ethanol tolerance. These strategies include the suppres-
sion of other fermentation products through lactate and acetate metabolic
pathway knockout in T. saccharolyticum JW/SL-YS485 and the overexpres-
sion of enzymes such as NAD(P)H-dependent ADH in T. mathranii, which
is directly related to increased ethanol production [109, 115]. The expression
dynamics of the ADH enzymes ADHA, ADHB, and ADHE, which have
a key role in ethanol formation in T. ethanolicus, were evaluated, and the
expression of these enzymes was observed to be affected at high ethanol
concentrations [116].
In conclusion, the search for more efficient and resistant microorgan-
isms to produce ethanol continues. Microorganisms such as S. cerevisiae
and Z. mobilis have been extensively studied due to their ability to produce
ethanol; knowledge of their genomes and the application of omic techniques
have considerably illuminated their mechanisms of gene regulation during
ethanol production. Emerging microorganisms such as K. marxianus and C.
thermocellum have also aroused interest for use in ethanol production due
to their outstanding metabolic and physiological characteristics. However,
KEYWORDS
Author Copy
• Clostridium thermocellum
• ethanol production
• genetic regulation
• Kluyveromyces marxianus
• Saccharomyces cerevisiae
• Zymomonas mobilis
REFERENCES
1. Prinz, B., & Lang, C., (2004). Gene regulation in yeast. Genet. Biotechnol., 129–145.
https://doi.org/10.1007/978-3-662-07426-8_8.
2. Barbosa De, M. F. S., & Ingram, L. O., (1994). Expression of the Zymomonas mobilis
alcohol dehydrogenase II (AdhB) and pyruvate decarboxylase (Pdc) genes in Bacillus.
Curr. Microbiol., 28(5), 279–282. https://doi.org/10.1007/BF01573206.
3. Xue, T., Liu, K., Chen, D., Yuan, X., Fang, J., Yan, H., Huang, L., Chen, Y., & He, W.,
(2018). Improved bioethanol production using CRISPR/Cas9 to disrupt the ADH2 gene
in Saccharomyces cerevisiae. World J. Microbiol. Biotechnol., 34(10), 154, 1–12. https://
doi.org/10.1007/s11274-018-2518-4
4. Lo, J., Zheng, T., Hon, S., Olson, D. G., & Lynd, L. R., (2015). The bifunctional alcohol
and aldehyde dehydrogenase gene, AdhE, Is necessary for ethanol production in
Clostridium thermocellum and Thermoanaerobacterium saccharolyticum. J. Bacteriol.,
197(8), 1386–1393. https://doi.org/10.1128/JB.02450-14.
5. Favaro, L., Jansen, T., & Van, Z. W. H., (2019). Exploring industrial and natural
Saccharomyces cerevisiae strains for the bio-based economy from biomass: The case
of bioethanol. Crit. Rev. Biotechnol., 39(6), 800–816. https://doi.org/10.1080/0738855
1.2019.1619157.
6. Yang, J., Ding, M. Z., Li, B. Z., Liu, Z. L., Wang, X., & Yuan, Y. J., (2012). Integrated
phospholipidomics and transcriptomics analysis of Saccharomyces cerevisiae with
enhanced tolerance to a mixture of acetic acid, furfural, and phenol. Omi. A J. Integr.
Biol., 16(7,8), 374–386. https://doi.org/10.1089/omi.2011.0127.
Apple Academic Press
7. Kim, D., & Hahn, J. S., (2013). Roles of the Yap1 transcription factor and antioxidants
in Saccharomyces cerevisiae’s tolerance to furfural and 5-hydroxymethylfurfural, which
function as thiol-reactive electrophiles generating oxidative stress. Appl. Environ.
Microbiol., 79(16), 5069–5077. https://doi.org/10.1128/AEM.00643-13.
8. Ishida, Y., Nguyen, T. T. M., & Izawa, S., (2017). The yeast ADH7 promoter enables
gene expression under pronounced translation repression caused by the combined stress
of vanillin, furfural, and 5-hydroxymethylfurfural. J. Biotechnol., 252, 65–72. https://
doi.org/10.1016/j.jbiotec.2017.04.024.
9. Liu, Z. L. L., (2018). Understanding the tolerance of the industrial yeast Saccharomyces
Author Copy
cerevisiae against a major class of toxic aldehyde compounds. Appl. Microbiol.
Biotechnol., 102(13), 5369–5390. https://doi.org/10.1007/s00253-018-8993-6.
10. Flores-Cosio, G., Arellano-Plaza, M., Gschaedler, A., & Amaya-Delgado, L., (2018).
Physiological response to furan derivatives stress by Kluyveromyces marxianus SLP1
in ethanol production. Rev. Mex. Ing. Quim., 17(1), 189–202. https://doi.org/10.24275/
uam/izt/dcbi/revmexingquim/2018v17n1/Flores.
11. Flores-Cosío, G., Herrera-López, E. J., Arellano-Plaza, M., Gschaedler-Mathis, A.,
Sanchez, A., & Amaya-Delgado, L., (2019). Dielectric property measurements as a
method to determine the physiological state of Kluyveromyces marxianus and Saccha-
romyces cerevisiae stressed with furan aldehydes. Appl. Microbiol. Biotechnol., 103(23,
24), 9633–9642. https://doi.org/10.1007/s00253-019-10152-2.
12. Guo, W., Chen, Y., Wei, N., & Feng, X., (2016). Investigate the metabolic reprogramming
of Saccharomyces cerevisiae for enhanced resistance to mixed fermentation inhibitors
via 13C metabolic flux analysis. PLoS One, 11(8), 1–15. https://doi.org/10.1371/journal.
pone.0161448.
13. Dellomonaco, C., Fava, F., & Gonzalez, R., (2010). The path to next generation biofuels:
Successes and challenges in the era of synthetic biology. Microb. Cell Fact., 9, 1–15.
https://doi.org/10.1186/1475-2859-9-1.
14. Matsushika, A., Inoue, H., Kodaki, T., & Sawayama, S., (2009). Ethanol production from
xylose in engineered Saccharomyces cerevisiae strains: Current state and perspectives.
Appl. Microbiol. Biotechnol., 84(1), 37–53. https://doi.org/10.1007/s00253-009-2101-x.
15. Wu, G., Xu, Z., & Jönsson, L. J., (2017). Profiling of Saccharomyces cerevisiae
transcription factors for engineering the resistance of yeast to lignocellulose-derived
inhibitors in biomass conversion. Microb. Cell Fact., 16(1), 1–15. https://doi.org/10.1186/
s12934-017-0811-9.
16. Greetham, D., Wimalasena, T. T., Leung, K., Marvin, M. E., Chandelia, Y., Hart, A. J.,
Phister, T. G., et al., (2014). The genetic basis of variation in clean lineages of Saccharomyces
cerevisiae in response to stresses encountered during bioethanol fermentations. PLoS One,
9(8), 1–14. https://doi.org/10.1371/journal.pone.0103233.
17. Cunha, J. T., Romaní, A., Costa, C. E., Sá-Correia, I., & Domingues, L., (2019). Molecular
and physiological basis of Saccharomyces cerevisiae tolerance to adverse lignocellulose-
based process conditions. Appl. Microbiol. Biotechnol., 103(1), 159–175. https://doi.org/
10.1007/s00253-018-9478-3.
18. Beckner, M., Ivey, M. L., & Phister, T. G., (2011). Microbial contamination of fuel
ethanol fermentations. Lett. Appl. Microbiol., 53(4), 387–394. https://doi.org/10.1111/j.
1472-765X.2011.03124.x.
19. Meijnen, J. P., Randazzo, P., Foulquié-Moreno, M. R., Van, D. B. J., Vandecruys, P.,
Apple Academic Press
Stojiljkovic, M., Dumortier, F., et al., (2016). Polygenic analysis and targeted improvement
of the complex trait of high acetic acid tolerance in the yeast Saccharomyces cerevisiae.
Biotechnol. Biofuels, 9(1), 1–18. https://doi.org/10.1186/s13068-015-0421-x.
20. Mira, N. P., Becker, J. D., & Sá-Correia, I., (2010). Genomic expression program
involving the Haa1p-regulon in Saccharomyces cerevisiae response to acetic acid. Omi.
A J. Integr. Biol., 14(5), 587–601. https://doi.org/10.1089/omi.2010.0048.
21. Mira, N. P., Teixeira, M. C., & Sá-Correia, I., (2010). Adaptive response and tolerance to
weak acids in Saccharomyces cerevisiae: A genome-wide view. Omi. A J. Integr. Biol.,
14(5), 525–540. https://doi.org/10.1089/omi.2010.0072.
Author Copy
22. Teixeira, M. C., Mira, N. P., & Sá-Correia, I., (2011). A genome-wide perspective on
the response and tolerance to food-relevant stresses in Saccharomyces cerevisiae. Curr.
Opin. Biotechnol., 22(2), 150–156. https://doi.org/10.1016/j.copbio.2010.10.011.
23. Saini, P., Beniwal, A., Kokkiligadda, A., & Vij, S., (2018). Response and tolerance of
yeast to changing environmental stress during ethanol fermentation. Process Biochem.,
72, 1–12. https://doi.org/10.1016/j.procbio.2018.07.001.
24. Almeida, J. R. M., Röder, A., Modig, T., Laadan, B., Lidén, G., & Gorwa-Grauslund,
M. F., (2008). NADH- vs NADPH-coupled reduction of 5-hydroxymethyl furfural
(HMF) and its implications on product distribution in Saccharomyces cerevisiae. Appl.
Microbiol. Biotechnol., 78(6), 939–945. https://doi.org/10.1007/s00253-008-1364-y.
25. Liu, Z. L., & Moon, J., (2009). A novel NADPH-dependent aldehyde reductase gene
from Saccharomyces cerevisiae NRRL Y-12632 involved in the detoxification of
aldehyde inhibitors derived from lignocellulosic biomass conversion. Gene, 446(1),
1–10. https://doi.org/10.1016/j.gene.2009.06.018.
26. Moon, J., & Liu, Z. L., (2012). Engineered NADH-dependent GRE2 from Saccharomyces
cerevisiae by directed enzyme evolution enhances HMF reduction using additional
cofactor NADPH. Enzyme Microb. Technol., 50(2), 115–120. https://doi.org/10.1016/j.
enzmictec.2011.10.007.
27. Larroy, C., Fernández, M. R., González, E., Parés, X., & Biosca, J. A., (2003). Properties
and functional significance of Saccharomyces cerevisiae ADHVI. Chem. Biol. Interact.,
143, 144, 229–238. https://doi.org/10.1016/S0009-2797(02)00166-7.
28. Adeboye, P. T., Bettiga, M., & Olsson, L., (2017). ALD5, PAD1, ATF1 and ATF2 facilitate
the catabolism of coniferyl aldehyde, ferulic acid and p-coumaric acid in Saccharomyces
cerevisiae. Sci. Rep., 7, 1–13. https://doi.org/10.1038/srep42635.
29. Ma, M., & Liu, Z. L., (2010). Comparative transcriptome profiling analyses during
the Lag phase uncover YAP1, PDR1, PDR3, RPN4, and HSF1 as key regulatory
genes in genomic adaptation to the lignocellulose derived inhibitor HMF for
Saccharomyces cerevisiae. BMC Genomics, 11(660), 1–19. http://www.biomedcentral.
com/1471-2164/11/660.
30. Wang, H., Li, Q., Zhang, Z., Zhou, C., Ayepa, E., Abrha, G. T., Han, X., et al., (2019).
YKL107W from Saccharomyces cerevisiae encodes a novel aldehyde reductase
for detoxification of acetaldehyde, glycolaldehyde, and furfural. Appl. Microbiol.
Biotechnol., 103(14), 5699–5713. https://doi.org/10.1007/s00253-019-09885-x.
31. Allen, S. A., Clark, W., McCaffery, J. M., Cai, Z., Lanctot, A., Slininger, P. J., Liu, Z. L.,
& Gorsich, S. W., (2010). Furfural induces reactive oxygen species accumulation and
cellular damage in Saccharomyces cerevisiae. Biotechnol. Biofuels, 3, 2.
32. Voulgaridou, G. P., Anestopoulos, I., Franco, R., Panayiotidis, M. I., & Pappa, A.,
Apple Academic Press
Author Copy
Saccharomyces cerevisiae for simultaneous saccharification and ethanol fermentation
from industrial waste corncob residues. Bioresour. Technol., 157, 6–13. https://doi.
org/10.1016/j.biortech.2014.01.060.
35. Wright, J., Bellissimi, E., De Hulster, E., Wagner, A., Pronk, J. T., & Van, M. A. J. A.,
(2011). Batch and Continuous culture-based selection strategies for acetic acid tolerance
in xylose-fermenting Saccharomyces cerevisiae. FEMS Yeast Res., 11(3), 299–306.
https://doi.org/10.1111/j.1567-1364.2011.00719.x.
36. Li, H., Ma, M. L., Luo, S., Zhang, R. M., Han, P., & Hu, W., (2012). Metabolic
responses to ethanol in Saccharomyces cerevisiae using a gas chromatography tandem
mass spectrometry-based metabolomics approach. Int. J. Biochem. Cell Biol., 44(7),
1087–1096. https://doi.org/10.1016/j.biocel.2012.03.017.
37. Nugroho, R. H., Yoshikawa, K., & Shimizu, H., (2015). Metabolomic analysis of acid
stress response in Saccharomyces cerevisiae. J. Biosci. Bioeng., 120(4), 396–404. https://
doi.org/10.1016/j.jbiosc.2015.02.011.
38. Leupold, S., Hubmann, G., Litsios, A., Meinema, A. C., Takhaveev, V., Papagiannakis,
A., Niebel, B., et al., (2019). Saccharomyces cerevisiae goes through distinct metabolic
phases during its replicative lifespan. Elife, 8, 1–19. https://doi.org/10.7554/eLife.41046.
39. Ohta, E., Nakayama, Y., Mukai, Y., Bamba, T., & Fukusaki, E., (2016). Metabolomic
approach for improving ethanol stress tolerance in Saccharomyces cerevisiae. J. Biosci.
Bioeng., 121(4), 399–405. https://doi.org/10.1016/j.jbiosc.2015.08.006.
40. Raimondi, S., Zanni, E., Amaretti, A., Palleschi, C., Uccelletti, D., & Rossi, M., (2013).
Thermal adaptability of Kluyveromyces marxianus in recombinant protein production.
Microb. Cell Fact., 12(1), 1–7. https://doi.org/10.1186/1475-2859-12-34.
41. Morrissey, J. P., Etschmann, M. M. W., Schrader, J., & De Billerbeck, G. M., (2015).
Cell factory applications of the yeast Kluyveromyces marxianus for the biotechnological
production of natural flavor and fragrance molecules. Yeast, (32), 3–16. https://doi.
org/10.1002/yea.
42. Iñiguez, M. L. E., Arellano-Plaza, M., Prado-Montes De, O. E., Kirchmayr, M. R.,
Segura-García, L. E., Amaya-Degado, L., & Gschaedler, M., (2019). The production of
esters and gene expression by Saccharomyces cerevisiae during fermentation on Agave
tequilana juice in continuous cultures. Rev. Mex. Ing. Química, 18(2), 451–462. https://
doi.org/10.24275/uam/izt/dcbi/revmexingquim/2019v18n2/iniguez.
43. Rodrussamee, N., Lertwattanasakul, N., Hirata, K., Suprayogi, Limtong, S., Kosaka,
T., & Yamada, M., (2011). Growth and ethanol fermentation ability on hexose and
pentose sugars and glucose effect under various conditions in thermotolerant yeast
Kluyveromyces marxianus. Appl. Microbiol. Biotechnol., 90(4), 1573–1586. https://doi.
org/10.1007/s00253-011-3218-2.
44. García-Aparicio, M. P., Oliva, J. M., Manzanares, P., Ballesteros, M., Ballesteros, I.,
Apple Academic Press
González, A., & Negro, M. J., (2011). Second-generation ethanol production from
steam exploded barley straw by Kluyveromyces marxianus CECT 10875. Fuel, 90(4),
1624–1630. https://doi.org/10.1016/j.fuel.2010.10.052.
45. Limtong, S., Sringiew, C., & Yongmanitchai, W., (2007). Production of fuel ethanol at
high temperature from sugar cane juice by a newly isolated Kluyveromyces marxianus.
Bioresour. Technol., 98(17), 3367–3374. https://doi.org/10.1016/j.biortech.2006.10.044.
46. Moreno, A. D., Ibarra, D., Fernández, J. L., & Ballesteros, M., (2012). Different laccase
detoxification strategies for ethanol production from lignocellulosic biomass by the
thermotolerant yeast Kluyveromyces marxianus CECT 10875. Bioresour. Technol., 106,
Author Copy
101–109. https://doi.org/10.1016/j.biortech.2011.11.108.
47. Moreno, A. D., Ibarra, D., Ballesteros, I., González, A., & Ballesteros, M., (2013).
Comparing cell viability and ethanol fermentation of the thermotolerant yeast Kluyvero-
myces marxianus and Saccharomyces cerevisiae on steam-exploded biomass treated with
laccase. Bioresour. Technol., 135, 239–245. https://doi.org/10.1016/j.biortech.2012.11.095.
48. Nitiyon, S., Keo-oudone, C., Murata, M., Lertwattanasakul, N., Limtong, S., Kosaka,
T., & Yamada, M., (2016). Efficient conversion of xylose to ethanol by stress-tolerant
Kluyveromyces marxianus BUNL-21. Springerplus, 5(1), 1–12. https://doi.org/10.1186/
s40064-016-1881-6.
49. Pilap, W., Thanonkeo, S., Klanrit, P., & Thanonkeo, P., (2018). The potential of the newly
isolated thermotolerant Kluyveromyces marxianus for high-temperature ethanol production
using sweet sorghum juice. 3 Biotech, 8(2). https://doi.org/10.1007/s13205-018-1161-y.
50. Saini, J. K., Agrawal, R., Satlewal, A., Saini, R., Gupta, R., Mathur, A., & Tuli, D.,
(2015). Second generation bioethanol production at high gravity of pilot-scale pretreated
wheat straw employing newly isolated thermotolerant yeast Kluyveromyces marxianus
DBTIOC-35. RSC Adv., 5(47), 37485–37494. https://doi.org/10.1039/c5ra05792b.
51. Signori, L., Passolunghi, S., Ruohonen, L., Porro, D., & Branduardi, P., (2014). Effect of
oxygenation and temperature on glucose-xylose fermentation in Kluyveromyces marxianus
CBS712 strain. Microb. Cell Fact., 13(1), 1–13. https://doi.org/10.1186/1475-2859-13-51.
52. Suryawati, L., Wilkins, M. R., Bellmer, D. D., Huhnke, R. L., Maness, N. O., & Banat,
I. M., (2008). Simultaneous Saccharification and fermentation of Kanlow switchgrass
pretreated by hydrothermolysis using Kluyveromyces marxianus IMB4. Biotechnol.
Bioeng., 101(5), 894–902. https://doi.org/10.1002/bit.21965.
53. Tomás-Pejó, E., Oliva, J. M., González, A., Ballesteros, I., & Ballesteros, M., (2009).
Bioethanol production from wheat straw by the thermotolerant yeast Kluyveromyces
marxianus CECT 10875 in a simultaneous saccharification and fermentation fed-batch
process. Fuel, 88(11), 2142–2147. https://doi.org/10.1016/j.fuel.2009.01.014.
54. Zhang, B., Zhang, J., Wang, D., Han, R., Ding, R., Gao, X., Sun, L., & Hong, J., (2016).
Simultaneous fermentation of glucose and xylose at elevated temperatures co-produces
ethanol and xylitol through overexpression of a xylose-specific transporter in engineered
Kluyveromyces marxianus. Bioresour. Technol., 216, 227–237. https://doi.org/10.1016/j.
biortech.2016.05.068.
55. Sandoval-Nuñez, D., Arellano-Plaza, M., Gschaedler, A., Arrizon, J., & Amaya-
Delgado, L., (2018). A comparative study of lignocellulosic ethanol productivities by
14177–14192. https://doi.org/10.1039/c8ra00335a.
57. Matsumoto, I., Arai, T., Nishimoto, Y., Leelavatcharamas, V., Furuta, M., & Kishida,
M., (2018). Thermotolerant yeast Kluyveromyces marxianus reveals more tolerance to
heat shock than the brewery yeast Saccharomyces cerevisiae. Biocontrol Sci., 23(3),
133–138. https://doi.org/10.4265/bio.23.133.
58. Mo, W., Wang, M., Zhan, R., Yu, Y., He, Y., & Lu, H., (2019). Kluyveromyces marxianus
developing ethanol tolerance during adaptive evolution with significant improvements
of multiple pathways. Biotechnol. Biofuels, 12(1), 1–15. https://doi.org/10.1186/
s13068-019-1393-z.
Author Copy
59. Lertwattanasakul, N., Kosaka, T., Hosoyama, A., Suzuki, Y., Rodrussamee, N., Matsutani,
M., Murata, M., et al., (2015). Genetic basis of the highly efficient yeast Kluyveromyces
marxianus: Complete genome sequence and transcriptome analyses. Biotechnol. Biofuels,
8(1), 1–14. https://doi.org/10.1186/s13068-015-0227-x.
60. Li, P., Fu, X., Zhang, L., Zhang, Z., Li, J., & Li, S., (2017). The transcription factors
Hsf1 and Msn2 of thermotolerant Kluyveromyces marxianus promote cell growth and
ethanol fermentation of Saccharomyces cerevisiae at high temperatures. Biotechnol.
Biofuels, 10(1), 1–13. https://doi.org/10.1186/s13068-017-0984-9.
61. Schabort, D. T. W. P., Kilian, S. G., & Du Preez, J. C., (2018). Gene regulation in
Kluyveromyces marxianus in the context of chromosomes. PLoS One, 13(1), 1–16.
https://doi.org/10.1371/journal.pone.0190913.
62. Gao, J., Feng, H., Yuan, W., Li, Y., Hou, S., Zhong, S., & Bai, F., (2017). Enhanced
fermentative performance under stresses of multiple lignocellulose-derived inhibitors
by overexpression of a typical 2-Cys peroxiredoxin from Kluyveromyces marxianus.
Biotechnol. Biofuels, 10(1), 1–13. https://doi.org/10.1186/s13068-017-0766-4.
63. Kwon, D. H., Park, J. B., Hong, E., & Ha, S. J., (2019). Ethanol production from xylose
is highly increased by the Kluyveromyces marxianus mutant 17694-DH1. Bioprocess
Biosyst. Eng., 42(1), 63–70. https://doi.org/10.1007/s00449-018-2014-0.
64. Zhang, J., Zhang, B., Wang, D., Gao, X., Sun, L., & Hong, J., (2015). Rapid ethanol
production at elevated temperatures by engineered thermotolerant Kluyveromyces
marxianus via the NADP(H)-preferring xylose reductase-xylitol dehydrogenase
pathway. Metab. Eng., 31, 140–152. https://doi.org/10.1016/j.ymben.2015.07.008.
65. Löbs, A. K., Engel, R., Schwartz, C., Flores, A., & Wheeldon, I., (2017). CRISPR-Cas9-
enabled genetic disruptions for understanding ethanol and ethyl acetate biosynthesis in
Kluyveromyces marxianus. Biotechnol. Biofuels, 10(1), 1–14. https://doi.org/10.1186/
s13068-017-0854-5.
66. Li, P., Fu, X., Li, S., & Zhang, L., (2018). Engineering TATA-binding protein Spt15
to improve ethanol tolerance and production in Kluyveromyces marxianus. Biotechnol.
Biofuels, 11(1), 1–13. https://doi.org/10.1186/s13068-018-1206-9.
67. Alvim, M. C. T., Vital, C. E., Barros, E., Vieira, N. M., Da Silveira, F. A., Balbino, T. R.,
Diniz, R. H. S., et al., (2019). Ethanol stress responses of Kluyveromyces marxianus CCT
7735 revealed by proteomic and metabolomic analyses. Antonie Van Leeuwenhoek, Int.
J. Gen. Mol. Microbiol., 112(6), 827–845. https://doi.org/10.1007/s10482-018-01214-y.
68. Lee, K. J., & Rogers, P. L., (1983). The fermentation kinetics of ethanol production by
Zymomonas mobilis. Chem. Eng. J., 27(2). https://doi.org/10.1016/0300-9467(83)80067-7.
69. Kremer, T. A., LaSarre, B., Posto, A. L., McKinlay, J. B., & Ingram, L. O., (2015). N2
gas is an effective fertilizer for bioethanol production by Zymomonas mobilis. Proc.
Natl. Acad. Sci. U. S. A., 112(7), 2222–2226. https://doi.org/10.1073/pnas.1420663112.
70. Wang, X., He, Q., Yang, Y., Wang, J., Haning, K., Hu, Y., Wu, B., et al., (2018). Advances
Apple Academic Press
and prospects in metabolic engineering of Zymomonas mobilis. Metab. Eng., 50, 57–73.
https://doi.org/10.1016/j.ymben.2018.04.001.
71. Yang, S., Fei, Q., Zhang, Y., Contreras, L. M., Utturkar, S. M., Brown, S. D., Himmel,
M. E., & Zhang, M., (2016). Zymomonas mobilis as a model system for production of
biofuels and biochemicals. Microb. Biotechnol., 9(6), 699–717. https://doi.org/10.1111/
1751-7915.12408.
72. Todhanakasem, T., Yodsanga, S., Sowatad, A., Kanokratana, P., Thanonkeo, P., & Champreda,
V., (2018). Inhibition analysis of inhibitors derived from lignocellulose pretreatment on the
metabolic activity of Zymomonas mobilis biofilm and planktonic cells and the proteomic
Author Copy
responses. Biotechnol. Bioeng., 115(1), 70–81. https://doi.org/10.1002/bit.26449.
73. He, M. X., Wu, B., Qin, H., Ruan, Z. Y., Tan, F. R., Wang, J. L., Shui, Z. X., et al.,
(2014). Zymomonas mobilis: A novel platform for future biorefineries. Biotechnol.
Biofuels, 7(1), 1–15. https://doi.org/10.1186/1754-6834-7-101.
74. Kannuchamy, S., Mukund, N., & Saleena, L. M., (2016). Genetic engineering of
Clostridium thermocellum DSM1313 for enhanced ethanol production. BMC Biotechnol.,
16(1), 1–6. https://doi.org/10.1186/s12896-016-0260-2.
75. Seo, J. S., Chong, H., Park, H. S., Yoon, K. O., Jung, C., Kim, J. J., Hong, J. H., et
al., (2005). The genome sequence of the ethanologenic bacterium Zymomonas mobilis
ZM4. Nat. Biotechnol., 23(1), 63–68. https://doi.org/10.1038/nbt1045.
76. Yang, S., Pappas, K. M., Hauser, L. J., Land, M. L., Chen, G. L., Hurst, G. B., Pan, C.,
et al., (2009). Improved genome annotation for Zymomonas mobilis. Nat. Biotechnol.,
27(10), 893–894. https://doi.org/10.1038/nbt1009-893.
77. Anggarini, S., Murata, M., Kido, K., Kosaka, T., Sootsuwan, K., Thanonkeo, P., &
Yamada, M., (2020). Improvement of thermotolerance of Zymomonas mobilis by
genes for reactive oxygen species-scavenging enzymes and heat shock proteins. Front.
Microbiol., 10, 1–14. https://doi.org/10.3389/fmicb.2019.03073.
78. Charoensuk, K., Irie, A., Lertwattanasakul, N., Sootsuwan, K., Thanonkeo, P., & Yamada,
M., (2011). Physiological Importance of cytochrome c peroxidase in ethanologenic
thermotolerant Zymomonas mobilis. J. Mol. Microbiol. Biotechnol., 20(2), 70–82.
https://doi.org/10.1159/000324675.
79. Yang, S., Tschaplinski, T. J., Engle, N. L., Carroll, S. L., Martin, S. L., Davison, B. H.,
Palumbo, A. V., et al., (2009). Transcriptomic and metabolomic profiling of Zymomonas
mobilis during aerobic and anaerobic fermentations. BMC Genomics, 10. https://doi.
org/10.1186/1471-2164-10-34.
80. Jönsson, L. J., & Martín, C., (2016). Pretreatment of lignocellulose: Formation of
inhibitory by-products and strategies for minimizing their effects. Bioresour. Technol.,
199, 103–112. https://doi.org/10.1016/j.biortech.2015.10.009.
81. Yang, S., Vera, J. M., Grass, J., Savvakis, G., Moskvin, O. V., Yang, Y., McIlwain, S. J.,
et al., (2018). Complete genome sequence and the expression pattern of plasmids of the
model ethanologen Zymomonas mobilis ZM4 and its xylose-utilizing derivatives 8b and
2032. Biotechnol. Biofuels, 11(1), 1–20. https://doi.org/10.1186/s13068-018-1116-x.
82. Yi, X., Gu, H., Gao, Q., Liu, Z. L., & Bao, J., (2015). Transcriptome analysis of Zymomonas
mobilis ZM4 reveals mechanisms of tolerance and detoxification of phenolic aldehyde
Author Copy
Biofuels, 5, 1–10. https://doi.org/10.1186/1754-6834-5-75.
87. Brás, J. L. A., Pinheiro, B. A., Cameron, K., Cuskin, F., Viegas, A., Najmudin, S., Bule,
P., et al., (2016). Diverse specificity of cellulosome attachment to the bacterial cell
surface. Sci. Rep., 6, 1–12. https://doi.org/10.1038/srep38292.
88. Singh, N., Mathur, A. S., Gupta, R. P., Barrow, C. J., Tuli, D., & Puri, M., (2018).
Enhanced cellulosic ethanol production via consolidated bioprocessing by Clostridium
thermocellum ATCC 31924. Bioresour. Technol., 250, 860–867. https://doi.org/10.1016/j.
biortech.2017.11.048.
89. Zhu, X., Cui, J., Feng, Y., Fa, Y., Zhang, J., & Cui, Q., (2013). Metabolic adaption of
ethanol-tolerant Clostridium thermocellum. PLoS One, 8(7), 1–9. https://doi.org/10.1371/
journal.pone.0070631.
90. Tian, L., Perot, S. J., Hon, S., Zhou, J., Liang, X., Bouvier, J. T., Guss, A. Met al., (2017).
Enhanced ethanol formation by Clostridium thermocellum via pyruvate decarboxylase.
Microb. Cell Fact., 16(1), 1–10. https://doi.org/10.1186/s12934-017-0783-9.
91. Holwerda, E. K., Olson, D. G., Ruppertsberger, N. M., Stevenson, D. M., Murphy, S. J. L.,
Maloney, M. I., Lanahan, A. A., et al., (2020). Metabolic and evolutionary responses of
Clostridium thermocellum to genetic interventions aimed at improving ethanol production.
Biotechnol. Biofuels, 13(1), 1–20. https://doi.org/10.1186/s13068-020-01680-5.
92. Xu, Q., Resch, M. G., Podkaminer, K., Yang, S., Baker, J. O., Donohoe, B. S., Wilson,
C., et al., (2016). Cell biology: Dramatic performance of Clostridium thermocellum
explained by its wide range of cellulase modalities. Sci. Adv., 2(2). https://doi.org/10.1126/
sciadv.1501254.
93. Newcomb, M., Chen, C. Y., & Wu, J. H. D., (2007). Induction of the CelC operon of
Clostridium thermocellum by laminaribiose. Proc. Natl. Acad. Sci. U. S. A, 104(10),
3747–3752. https://doi.org/10.1073/pnas.0700087104.
94. Newcomb, M., Millen, J., Chen, C. Y., & Wu, J. H. D., (2011). Co-transcription of the
CelC gene cluster in Clostridium thermocellum. Appl. Microbiol. Biotechnol., 90(2),
625–634. https://doi.org/10.1007/s00253–011–3121-x.
95. Wilson, C. M., Klingeman, D. M, Schlachter, C., Syed, M. H., Wu, C., Guss, A. M., Brown, S.
D., (2017). LacI transcriptional regulatory networks in Clostridium thermocellum DSM1313.
Appl Environ Microbiol., 83, e02751–16. https://doi.org/10.1128/AEM.02751-16.
96. Choi, J., Klingeman, D. M., Brown, S. D., & Cox, C. D., (2017). The LacI family
protein GlyR3 co-regulates the CelC operon and ManB in Clostridium thermocellum.
Biotechnol. Biofuels, 10(1), 1–11. https://doi.org/10.1186/s13068-017-0849-2.
97. Nataf, Y., Bahari, L., Kahel-Raifer, H., Borovok, I., Lamed, R., Bayer, E. A., Sonenshein,
A. L., & Shoham, Y., (2010). Clostridium thermocellum cellulosomal genes are regulated
by extracytoplasmic polysaccharides via alternative sigma factors. Proc. Natl. Acad.
Sci. U. S. A., 107(43), 18646–18651. https://doi.org/10.1073/pnas.1012175107.
Apple Academic Press
98. Ortiz De, O. L., Lamed, R., Liu, Y. J., Xu, J., Cui, Q., Feng, Y., Shoham, Y., et al., (2018).
Regulation of biomass degradation by alternative σ factors in cellulolytic clostridia. Sci.
Rep., 8(1), 1–11. https://doi.org/10.1038/s41598-018-29245-5.
99. Nataf, Y., Yaron, S., Stahl, F., Lamed, R., Bayer, E. A., Scheper, T. H., Sonenshein, A. L.,
& Shoham, Y., (2009). Cellodextrin and laminaribiose ABC transporters in Clostridium
thermocellum. J. Bacteriol., 91(1), 203–209. https://doi.org/10.1128/JB.01190-08.
100. Olson, D. G., Maloney, M., Lanahan, A. A., Hon, S., Hauser, L. J., & Lynd, L. R., (2015).
Identifying promoters for gene expression in Clostridium thermocellum. Metab. Eng.
Commun., 2, 23–29. https://doi.org/10.1016/j.meteno.2015.03.002.
Author Copy
101. Tian, L., Cervenka, N. D., Low, A. M., Olson, D. G., & Lynd, L. R., (2019). A mutation
in the AdhE alcohol dehydrogenase of Clostridium thermocellum Increases tolerance to
several primary alcohols, including isobutanol, n-butanol and ethanol. Sci. Rep., 9(1),
1–7. https://doi.org/10.1038/s41598-018-37979-5.
102. Williams, T. I., Combs, J. C., Lynn, B. C., & Strobel, H. J., (2007). Proteomic profile
changes in membranes of ethanol-tolerant Clostridium thermocellum. Appl. Microbiol.
Biotechnol., 74(2), 422–432. https://doi.org/10.1007/s00253-006-0689-7.
103. Shao, X., Raman, B., Zhu, M., Mielenz, J. R., Brown, S. D., Guss, A. M., & Lynd, L. R.,
(2011). Mutant selection and phenotypic and genetic characterization of ethanol-tolerant
strains of Clostridium thermocellum. Appl. Microbiol. Biotechnol., 92(3), 641–652.
https://doi.org/10.1007/s00253-011-3492-z.
104. Cui, J., Stevenson, D., Korosh, T., Amador-Noguez, D., Olson, D. G., & Lynd, L. R., (2020).
Developing a cell-free extract reaction (CFER) system in Clostridium thermocellum to
identify metabolic limitations to ethanol production. Front. Energy Res., 8. https://doi.
org/10.3389/fenrg.2020.00072.
105. Stevenson, D. M., & Weimer, P. J., (2005). Expression of 17 genes in. Society, 71(8),
4672–4678. https://doi.org/10.1128/AEM.71.8.4672.
106. Riederer, A., Takasuka, T. E., Makino, S. I., Stevenson, D. M., Bukhman, Y. V.,
Elsen, N. L., & Fox, B. G., (2011). Global gene expression patterns in Clostridium
thermocellum as determined by microarray analysis of chemostat cultures on cellulose
or cellobiose. Appl. Environ. Microbiol., 77(4), 1243–1253. https://doi.org/10.1128/
AEM.02008-10.
107. Wilson, C. M., Rodriguez, M., Johnson, C. M., Martin, S. L., Chu, T. M., Wolfinger,
R. D., Hauser, L. J., et al., (2013). Global transcriptome analysis of Clostridium
thermocellum ATCC 27405 during growth on dilute acid pretreated Populus and
switchgrass. Biotechnol. Biofuels, 6(1), 1–18. https://doi.org/10.1186/1754–6834-6-179.
108. Raman, B., McKeown, C. K., Rodriguez, M., Brown, S. D., & Mielenz, J. R., (2011).
Transcriptomic analysis of Clostridium thermocellum ATCC 27405 cellulose fermenta-
tion. BMC Microbiol., 11. https://doi.org/10.1186/1471-2180-11-134.
109. Yao, S., & Mikkelsen, M. J., (2010). Metabolic engineering to improve ethanol production
in Thermoanaerobacter mathranii. Appl. Microbiol. Biotechnol., 88(1), 199–208. https://
doi.org/10.1007/s00253-010-2703-3.
110. Xue, Y., Xu, Y., Liu, Y., Ma, Y., & Zhou, P., (2001). Thermoanaerobacter tengcongensis
Sp. Nov., a novel anaerobic, saccharolytic, thermophilic bacterium isolated from a hot
spring in tengcong, China. Int. J. Syst. Evol. Microbiol., 51(4), 1335–1341. https://doi.
org/10.1099/00207713-51-4-1335.
111. Slobodkin, A. I., Tourova, T. P., Kuznetsov, B. B., Kostrikina, N. A., Chernyh, N. A.,
& Bonch-Osmolovskaya, E. A., (1999). Thermoanaerobacter siderophilus sp. Nov., a
Apple Academic Press
Author Copy
pentosaceus at 70°C in an up-flow anaerobic sludge blanket (UASB) reactor. Bioresour.
Technol., 143, 598–607. https://doi.org/10.1016/j.biortech.2013.06.056.
114. Michael, S. S., & Orlygsson, J., (2019). Progress in second generation ethanol production
with thermophilic bacteria. Fuel Ethanol Prod. from Sugarcane. https://doi.org/10.5772/
intechopen.78020.
115. Desai, S. G., Guerinot, M. L., & Lynd, L. R., (2004). Cloning of L-lactate dehydrogenase
and elimination of lactic acid production via gene knockout in Thermoanaerobacterium
saccharolyticum JW/SL-YS485. Appl. Microbiol. Biotechnol., 65(5), 600–605. https://
doi.org/10.1007/s00253-004-1575-9.
116. Pei, J., Zhou, Q., Jiang, Y., Le, Y., Li, H., Shao, W., & Wiegel, J., (2010). Thermoanaero-
bacter Spp. control ethanol pathway via transcriptional regulation and versatility of key
enzymes. Metab. Eng., 12(5), 420–428. https://doi.org/10.1016/j.ymben.2010.06.001.
Author Copy
S. SÁNCHEZ-MUÑOZ,1 M. J. CASTRO-ALONSO,1 F. G. BARBOSA,1
E. MIER-ALBA,1 T. R. BALBINO,1 D. RUBIO-RIBEAUX,1
I. O. HERNÁNDEZ-DE LIRA,2 J. C. SANTOS,3 C. N. AGUILAR,4 and
S. S. DA SILVA1
1
Bioprocesses and Sustainable Products Laboratory,
Department of Biotechnology, Engineering School of Lorena,
University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil,
E-mail: silviosilverio@gmail.com (S. S. Da Silva)
2
Bioremediation Laboratory, Biological Science Faculty,
Autonomous University of Coahuila (UAdeC), Torreón Campus,
Coahuila–27276, México
Bioprocesses, Biopolymers, Simulation, and Modeling Laboratory,
3
ABSTRACT
The price fluctuation of petroleum-based fuels makes biofuels one of the most
promising alternative energies for global economies. Since the biological
production of fuels, from vegetal biomass, offers sustainable and economically
attractive options compared to petroleum-based production. However, the
conversion to fuel. Those steps are being consolidated into single microbial
processes using metabolically engineered species (e.g., Saccharomyces
cerevisiae, Zimmomona mobilis, and Clostridium thermocellum). This chapter
will review the advance in metabolic engineering and synthetic biology on
the main ethanol producers’ microorganisms, which allow the development
of new engineered systems aiming for a better transition from fossil fuels to
biofuels.
Author Copy
5.1 INTRODUCTION
databases, growth speed, low incubation costs, and their now established use
in industrial bioprocesses [6, 11, 12]. However, other microorganisms such
as the bacterium Zymomonas mobilis and the fungus Clostridium thermo-
cellum are of great interest in research field, because of their versatility to
produce several biomolecules under industrial conditions, that made them
excellent alternatives for biotechnology [13, 14].
In this chapter, we have addressed the current state of metabolic engineering
Author Copy
in biofuel production and described the recent progress made in producing
bioethanol from the main genetically modified microorganisms.
Author Copy
process of cellular components; in some cases, its interruption enhance the
biomass growth rate (ex. yeast) and consequently, a positive influence in the
production of some products like ethanol could occur [27–29]. Fermentation
has been well-studied for the application of several molecular tools. The most
used one could be the overexpression of specific enzymes. For example, in
citric acid production, pyruvate carboxylase is a key enzyme in the reduction
of the tricarboxylic acid cycle, and it is closely related to the formation of this
important biomolecule, thus the overexpression of this enzyme could lead to
a higher production of this organic acid [30, 31]. Other examples of metabolic
engineering tools applied to microorganisms for enhancing the production of
biomolecules are shown in Table 5.1.
Ethanol Saccharomyces cerevisiae Disruption of ATG32 region 2.76% Beverages (sake) [29]
(Y)
Cellulases/ Trichoderma reesei RUT Construction of transcriptional 50–80% Saccharification of [35]
Xylanases C30 (F) activation linked to a domain of herpes lignocellulosic biomass
simplex virus protein VP16
Pullulan Aureobasidium pullulans Homologous expression of UGPase 1.7-fold Blood plasma substitutes, [36]
NRRL Y-12974 (F) food preservation,
adhesive, and others
Author Copy
102 Bioethanol: Biochemistry and Biotechnological Advances
One of the major challenges for achieve high levels of biofuels production
Author Copy
is that fermentation performance is affected by toxicity of end products [43].
The toxicity effects of biofuels, particularly organic solvents, are related
to the logarithm of its octanol-water partition coefficient (logP). Organic
solvents with logP values between a range of 0.8 and 5 cause disruption
of cellular membranes and interferes with several vital metabolic functions,
such as membrane transport and energy generation. Moreover, solvents
generate damage to biological macromolecules, such as DNA and RNA [44].
Several genetic engineering strategies had been employed to attenuate the
toxicity of organic solvents and biofuels, and to elucidate stress responses of
microorganism through the overexpression of heat shock protein genes [45],
redirection of carbon metabolic flux [46], regulation extrusion of solvents
efflux pumps genes [47] among others.
HSP are involved in the synthesis, transport, and folding of proteins closely
related to metabolic activities [43, 48]. The role of HSP in enhancement of
biofuel tolerance was showed for the first time in Clostridrium acetobutylicum.
In that study, the overexpression of chaperone GroESL improved tolerance to
n-butanol (85%) [45]. Another work from this research group demonstrated
that the overexpression of GroESL, also increased the expression of other
HSP involved in solvent tolerance, including dnaKJ, hsp18, hsp90 [49]. In
other studies, with isobutanol, the microbial transcriptional profiles revealed
main genes related to heat shock stress and protein misfolding, including
rpoH, dnaJ, htpG, and ibpAB [50, 51]. The same strategy was applied in
Escherichia coli, Lactobacillus plantarum and Zymomonas mobilis and
showed significant improvement in survival and viability of cells in presence
of ethanol, 1-butanol, and 2,3-butanediol by the overexpression of GroESL,
ClpB chaperones and SecB multi-tasking chaperone [44, 52–55]. On the
other hand, Zhang et al. [56] showed that genes involved in central carbon
flux and TCA cycle (gapA and sdhB) are closed related to end product
tolerance. Genes gapA (glyceraldehyde-3-phosphate dehydrogenase A) and
Author Copy
the moment, it is known that yjjZ gene plays a regulatory role in cellular
metabolism as small RNA (sRNA) [58–60]. In another work, Huang and
Geng [61] developed a novel replicative CRISPR/Cas9 plasmids system
(pdC99 and pPC9) to integrate the 2,3-butanediol pathway in Saccharomyces
cerevisiae that redirected carbon metabolic flux to improve tolerance and
lead to higher production of 2,3 butanediol (50.5 g/L).
Efflux pumps are membrane transporters responsible for recognizing
and exporting toxic compounds from the cytoplasm to out of cell. Those
transporters are comprised by three types of proteins: inner membrane
proteins, periplasmic linkers, and outer membrane channels [43]. Several
studies observed that overexpression of exogenous efflux pumps and the
cloning of their operons are effective strategies to enhance the tolerance
to biofuels. Early studies reported that genes encoding the efflux pump
SrpABC of Pseudomonas putida from the RND family were expressed in
the presence of n-butanol [62, 63]. Since the elucidation of this mechanism,
other studies revealed that the overexpression of efflux pump srpABC or
the srpB subunit from P. putida in E. coli enhances n-butanol tolerance
(20–40%) [64, 65]. Similarly, the co-overexpression of the membrane ATP
binding cassette ABC and the efflux pumps Snp2 and Pdr5 in S. cerevisiae
increased the tolerance and cellular growth when it was underexposed to
exogenous n-decane [66]. Reyes et al. [47], found that overexpression of
another membrane transporter associated to the genes ygfO, setA, mdtA, and
pgsA improve n-butanol tolerance in E. coli. Furthermore, this study also
exhibited that underexposure to n-butanol can lead to adaptive structural
changes of the membrane lipid structure, since pgsA gene is involved in
cardiolipin biosynthesis. In a recent work, a negative feedback network
was introduced to E. coli strain using promoter PgntK, which regulated the
expression of the butanol efflux pump AcrB and controls membrane protein
expression to optimize growth. That study reported an increase in tolerance
acrB gene in E. coli also raised tolerance of n-butanol in terms of cell density,
which was improved by 82.8%.
Author Copy
Inhibitors derived from fermentation process due to the pre-processing of
lignocellulosic material limits the metabolism of strains in terms of micro-
bial growth and biofuel production [71]. These inhibitors can be categorized
into three major types: (i) furan derivatives, (ii) short-chain aliphatic acids
and (iii) phenolic compounds, as reviewed by Palmqvist and Hahn-Hägerdal
[72]. Among these compounds, furan derivatives such as furfural and
5-hydroxymethylfurfural (HMF) are some of the most toxic substances
since they induced serious damages to cellular metabolism. Furfural and
HMF inhibits glycolytic and fermentative enzymes (pyruvate, acetaldehyde,
and alcohol dehydrogenases (ADHs)) and cause breaks in double-stranded
DNA [73]. Thus, furan derivates also generate reduction in concentration
of intracellular ATP and NAD(P)H because the energy flux is redirected to
pentose phosphate pathway (PPP) for repairing the induced damages [74,
75]. Moreover, furfural also affect mitochondria and vacuole membranes, as
it induces the formation of intracellular ROS [76]. All the cellular damages
mentioned above result in prolongation of lag phase of microbial growth and
decrease specific growth rate, which lead to low yield of biofuel produc-
tion [46]. In addition, furfural also increase the toxicity of acetic acid and
phenolic compounds [77, 78].
Extensive studies have been performed to improved microbial tolerance to
these inhibitors through classical and novel mutagenesis strategies. Miller et
al. [75] reported that silencing of NADPH-dependent oxidoreductases (yqhD
and dkgA) increased furfural tolerance in E. coli. In contrast, several recent
studies showed that overexpression of oxidoreductases and proteins involved
in oxidative stresses improved inhibitor tolerance. Co-overexpression of
oxidoreductases (yqhD and FucO) showed increased tolerance to furfural,
improved glucose utilization, and enhance the production of isobutanol
(110%) [79]. Similarly, Song et al. [80] employed the strategy of increasing
NAD(P)H pool through overexpression of pncB and nadE genes. These genes
are involved in nicotine amide salvage pathway that led to reduce furfural
toxicity and improved the production of isobutanol (2.5-fold higher). Oh et
al. [81] found that overexpression RCK1 gene encoding for a protein kinase
involved in oxidative stress improved acetic acid tolerance under glucose
Apple Academic Press
and xylose conditions and reduce 40% of intracellular ROS levels. More
recently, Liu et al. [82] showed that the manipulation of intracellular redox
potential of genes encoding cofactors related oxidoreductases in Z. mobilis
is a promising novel strategy to improve tolerance to acetic acid, furfural,
and phenolic compounds in industrial biofuels production. In other studies,
glycerol supplementation results in other promising strategy to furfural
detoxification [83, 84]. Agu et al. [84] demonstrated that overexpression
Author Copy
of two glycerol dehydrogenases dhaD1 and gldA1 reduced 68% of furfural
toxicity in Clostridium beijerinckii.
Author Copy
dynamically regulates the altruistic sacrifice of individual cells to reduce
metabolic heat release by detection of temperature and cell density. This
study showed that the synthetic IMHeRE system improves cell growth (10%)
at >40°C. Xu et al. [95] engineered an artificial antioxidant defense system
to improve thermo-tolerance of yeast. They introduced several antioxidant
genes from S. cerevisiae and Thermus thermophiles HB8 to construct “Angel
yeast.” This synthetic yeast showed an increase of thermotolerance in terms
of cell density (65.2%) at >40°C. Li et al. [87] used CRISPR/Cas-based
gene activation screening in S. cerevisiae and identified the essential role
of delta-9 desaturase gene OLE1 in increasing fatty acid unsaturation and
reduction of lipid peroxidation caused by heat stress at 42°C. Similarly, Li
et al. [96] elucidated the role of 4-Hydroxy-2,2,6,6-tetramethylpiperidin-
1-oxyl (TEMPOL) as a stress-alleviating agent in S. cerevisiae. TEMPOL is
a redox-cycling nitroxide and membrane-permeable antioxidant escorted by
dual redox potential forms. This system enhanced expression of important
genes and the activity of enzymatic antioxidant defense system, altering the
ratio of cofactors to improve the non-enzymatic antioxidant defense system,
or by directly degrading ROS with the help of NADH. Recently, Tao et al.
[97] used ‘Cas9 nickase-based genome editing’ and found that the inactiva-
tion of the gene MspI improved the thermotolerance of C. cellulolyticum.
Despite the MspI gene belongs to the restriction-modification system, this
study revealed that the disruption of MspI gene can influence the expression
of other thermotolerant genes.
on the source and species of plants [98]. Cellulose and hemicellulose are
polymers composed of hexoses (glucose, mannose, and galactose), pentoses
(arabinose and xylose), and sugar acids (uronic acids). Lignin is an amor-
phous macromolecule, composed by phenylpropanoids (guacyl and siringyl
Apple Academic Press
Author Copy
Hydrolysates obtained from sugar fractions of biomass are a mixture of
hexoses, pentoses, and products derived from lignin. The products derived
from lignin generate seriously microbial damages that inhibit fermentation
performance. The artificial modification and genetic engineering of plants
that result in low content of lignin are potential strategies to enhance plant
biomass hydrolysis to obtain maximum sugar releases [100]. Furthermore,
genetic strategies have more advantages than physical and chemical pretreat-
ments considering that they do not require additional energy or chemicals
input and result in lower environmental pollution [101].
The sense, antisense, or RNAi approaches has been used to upregula-
tion or downregulation of genes involved in lignin biosynthesis pathway to
alter their content, location, and type (Syringyl: Guaiacyl ratio) [102, 103].
Early studies revealed that the expression of p-coumaroyl-CoA 3-hydroxy-
lase (C3′H) gene is related to lignin synthesis [104–107]. Subsequently,
Coleman et al. [108] reported that RNAi suppression of C3′H expression in
hybrid poplar led to significant reduction of lignin content (56–59%). Simi-
larly, Chen and Dixon [109] reported that downregulation of the hydroxy-
cinnamoyl transferase (HCT) and caffeic acid 3-O-methyltransferase
(COMT) genes, reduced lignin content (40%) and increase sugar releases
(166%) in alfalfa. Shafrin et al. [110] introduced hpRNA-based vectors for
downregulation of Cinnamate 4-hydroxylase (C4H) and COMT genes in
jute, which showed significant reduction of soluble lignin (13–23%) and
fiber lignin (13–17%). RNAi downregulation of COMT gene also reduces
the Syringyl:Guaiacyl (S:G) ratio, improved the fermentable sugars yield
(>34%) and ethanol production yield (≥ 38%) in switchgrass and sugarcane
bagasse, respectively [111, 228]. In contrast, other studies showed that
reduction of S:G lignin ratio had no correlation with the reduction of biomass
recalcitrance [112, 113]. More recently, Zhang et al. [114] demonstrated
Author Copy
5.3.5 GENETIC ENGINEERING STRATEGIES TO IMPROVE
BIOETHANOL PRODUCTION
FIGURE 5.1 Bioethanol production from 1st and 2nd generation biomass: metabolic routes integration. Genes in the dotted box are the
most manipulated to improve metabolic performance for bioethanol production in Saccharomyces cerevisiae (•), Zymomonas mobilis (•) and
Author Copy
110 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
bioethanol production refineries [119, 120]. During fermentation, some
stress conditions, such as high temperatures, ethanol concentration and the
presence of sugars (e.g., xylose), can limit the process. These factors lead
to the inhibition of the production process and consequently, the reduction
of the ethanol yield [117, 121, 122]. Thus, recombinant DNA techniques
and genetic modifications are constantly carried out in yeasts, in order to
improve the performance of this microorganism [120, 122].
ethanol stress for yeast cells. Techniques such as global transcription machinery
engineering (gTME), are also tools for obtaining ethanol tolerant strains. In
this case, genes of mutant transcription factors (TFs) that control the genes of
cellular metabolism are introduced. Studies have shown that a recombinant
S. cerevisiae with a mutation in the SPT15 gene (gene encoding the TATA
binding factor) can exhibit a new pattern of gene expression for several genes.
This recombinant isolate, with a completely different gene expression pattern,
Author Copy
was able to tolerate high levels of ethanol [134].
Author Copy
Pentose fermentation difficulties have been solved using genetically modi-
fied organisms or co-culture of two yeast strains [152, 153]. To metabolize
xylose, the microorganism must be able to incorporate pentoses into the cell
via membrane transporters. Subsequently, two different pathways for xylose
isomerization in xylulose can be followed: the balanced redox oxidoreductase
and the isomerase pathway. The first occurs when the enzyme xylose reductase
(XR) converts D-xylose to xylitol, then, the enzyme xylitol dehydrogenase
(XD) transforms xylitol into D-xylulose. Later, xylose is isomerized to xylulose
by the enzyme Xylose Isomerase (XI) without the need for a cofactor [149,
154]. Finally, this D-xylulose is further phosphorylated by a xylulokinase in
D-xylulose-5-P, which can enter the PPP [232].
Studies have been successful in the production of ethanol from xylose
using recombinant S. cerevisiae. These strains are constructed by inserting
genes that encode the heterologous metabolic enzymes XR and xylitol
dehydrogenase (XDH) from Scheffersomyces stipitis, Candida intermedia or
other strains [155, 227, 233]. In addition to the construction of recombinant
yeasts, studies carried out genetic modifications to increase xylose consump-
tion. For example, the overexpression of the non-oxidative enzymes of the
PPP, endogenous XKS1 gene (encoding xylulokinase) and the deletion of
GRE3 (encoding reductase enzyme capable of converting xylose to xylitol
using NADPH) in S. cerevisiae, resulted in higher growth rate during xylose
fermentation [156]. Examples of other genetic strategies of yeasts to improve
bioethanol production are shown in Table 5.2.
Author Copy
TABLE 5.2
Apple Academic Press
(Continued) 114
Challenge Strategy Genetic Alteration Enhancement in Mutants Ethanol References
Yield (-fold)
Tolerance to Phenotypic analysis Simultaneous knockout of LEU4 and LEU9 Increased ethanol tolerance – [164]
ethanol and knockout (leading to the accumulation of valine) or
INM1 and INM2 (leading to the reduction
of inositol)
Xylose overexpression, Insertion of genes involved in the Increased capacity to ~2.0 [165]
assimilation evolutionary consumption of xylose (XYL1; XYL2; co-ferment glucose and xylose
engineering XKS1; TAL1; PYK1; MGT05196)
Author Copy
Metabolic Engineering of Yeast, Zymomonas mobilis 115
Author Copy
tors from lignocellulosic hydrolysates [13, 172]. In face of these limitations,
researchers have been focused on the development of genetic strategies in Z.
mobilis for improving ethanol production, as shown in Table 5.3.
As described before, some of the problems with a strong impact in ethanol
production by Z. mobilis are the inability of these bacteria to grow in presence
of inhibitors formed during the biomass pretreatment, such as acetic acid and
furfural [173]. The acetic acid, generated by hemicellulose and lignin deacety-
lation, can cross the cytoplasmic membrane of microorganisms by diffusion
when it is present in its undissociated form [174]. Inside the cell, acetic acid
is dissociated into proton and corresponding ion, resulting in a decrease of
intracellular pH, higher necessity of ATP to pumping protons out of the cells,
and growth inhibition [175, 176]. In an attempt to solve the problem caused
by the presence of acetic acid, Liu et al. [177] obtained acid-tolerant mutants
through chemical mutation and adaptative evolution. Similarly, Wu et al.
[178] studied acetic acid-tolerant mutant strains obtained via a multi-round
atmospheric and room temperature plasma (mARTP) mutagenesis.
On the other hand, furfural is formed during lignocellulose pretreat-
ment due to the dehydration of pentose sugars [166, 179]. Even in small
concentrations, this inhibitor can demonstrate a negative effect on terpenoid
biosynthesis and mRNAs related to the ED pathway, alterations of the central
carbon metabolism, membrane perturbation and decrease of NADH and ATP
concentrations [180, 181]. Regards to furfural inhibition, Z. mobilis mutants
were constructed by the error-prone PCR-based whole genome shuffling to
improve the tolerance to furfural [168]. The authors hypothesized that the
enhanced NADH-dependent furfural reductase activity detected during the
early log phase may be related to an increase of furfural tolerance of the
mutants, allowing an accelerated furfural detoxification process. Interest-
ingly, Wang et al. [182] considered both acetic acid and furfural inhibition
Author Copy
TABLE 5.3
Apple Academic Press
(Continued)
Phenolic Transformation Expression of NAD+-dependent aldehyde Increase of the conversion 1.77 [189]
aldehydes dehydratase rate of furfural, hydroxy-
bioconversion benzaldehyde, and vanillin
to less-toxic compounds
Use of biogas Atmospheric and room Nucleotides deletion in CDS of gene Ethanol was produced from 1.62 [195]
slurry such as temperature plasma ZMO_RS07255 (carbamoyl-phosphate biogas slurry to replace
substrate mutagenesis combined synthase large subunit) and between gene water and nutrients, under
with adaptive ZMO_RS06410 (FUSC family protein) sterilized and unsterilized
laboratory evolution and ZMO_RS06415 (DNA polymerase III condition.
Author Copy
118 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
acid synthesis mutation. Despite the improvement in ethanol tolerance, most
of the ethanol-tolerant mutant strains did not increase ethanol production.
However, Tan et al. [186] obtains ethanol-tolerant mutants and enhance
ethanol production yield by utilizing random mutagenesis of the sigma factor
RpoD protein (σ70).
Phenolic aldehydes are also produced during lignocellulose pretreatment
by the degradation of lignin components and can pass across cell membranes
due to its low molecular weight [166]. These inhibitors can cause damage to
internal structures and DNA, leading to the inhibition of RNA and protein
synthesis. Besides, it causes an increase of the cell membrane fluidity,
decrease of the intracellular potassium levels and cell growth, and cell
morphology abnormalities [187]. An alternative to reduce toxic effects of
these inhibitors is the direct conversion of phenolic aldehydes to less-toxic
compounds. However, the phenolic aldehydes are recalcitrant molecules
with low water solubility, which disturbs their bioconversion [188]. In this
context, Yi et al. [189] constructed an intracellular oxidative pathway to
enable simultaneous biodetoxification of phenolic aldehydes and fermen-
tation in Z. mobilis. Experiments in vivo demonstrated a strong oxidative
capacity of aldehyde dehydrogenase and upregulation of key genes of the
ED pathway and the oxidative phosphorylation.
Z. mobilis is a promising candidate to be used in the industrial process
with economically viable production [173]. Comparing to the model yeast
S. cerevisiae, this bacterium presents higher ethanol productivity (about
four-fold faster) [190, 191]. Moreover, the possibility to use low pH and
non-sterile conditions to grow Z. mobilis reduces the cost process and the
chance of contamination [192]. Therefore, the characteristics of Z. mobilis
added to studies of metabolic engineering that promote enhancement in
bioproduction, allow large-scale commercial production of bioethanol with
a positive economic balance for industries.
in the industrial field due to their potential for bioethanol production from a
wide range of carbohydrates (e.g., lignocellulosic biomass (LCB)) [196–199].
On this matter, Clostridium thermocellum is a gram-positive anaerobic
bacterium that grows at temperatures above 50°C and can metabolize different
substrates such as lignocellulose, cellobiose, xylose, and hemicelluloses [14,
200]. In the case of the lignocellulosic material, the hydrolysis occurs through
the cellulosome, which is a hydrolytic enzyme complex with a synergistic
Author Copy
action bound to a backbone scaffold protein and attached on the surface
of the cell wall of the microorganism [14]. Owing to C. thermocellum can
solubilize lignocellulose and produce ethanol as an additional fermentation
product, its application is considered as a low-cost consolidated bioprocess
with many advantages in one single step [201]. However, C. thermocellum is
inhibited in high concentrations of ethanol. This fact is related to alterations
in the membrane composition observed through the accumulation of long-
chain branched fatty acids during its culture [202]. Consequently, this ethanol
inhibition difficulties its widespread industrial adoption. In this sense,
metabolic engineering has been crucial to enhance the bioethanol production
capacities of C. thermocellum (Table 5.4). Hence, to turn this microorganism
a better ethanol producer, different protocols of transformation and deletion
have been widely tested by researchers [203–205]. For instance, deletion
experiments are routinely performed in strain DSM 1313 due to high
transformation efficiency [206, 207]. On the other hand, the development
of counter selectable markers such as tdk, hpt, and pyrF make possible an
easier manipulation of the C. thermocellum chromosome [203, 208, 209].
Furthermore, genetic tools such as gene deletion and targeted mutagenesis
have contributed with additional modifications as an attempt to meet the
industry demands [210, 211].
Among genetic tools, gene overexpression has provided substantial infor-
mation about the complexity of the clostridial metabolism and the possible
ways for increasing bioethanol production [212]. Similarly, heterologous
expression of bioethanol production pathways has been performed in native
strains as a potential solution to the commercialization of this biofuel [213,
214]. In addition, the carbon flux and metabolic pathways affected in C.
thermocellum by the inactivation or complete removal of specific genes have
been another alternative selected by researchers [215]. Since bioethanol
Author Copy
TABLE 5.4
Apple Academic Press
(Continued)
Increasing Cloning and Heterologous expression of pyruvate Two-fold increase in pyruvate 2.03 [222]
the affinity of transformation decarboxylase gene (pNW33N-pdc) carboxylase activity
the pyruvate from Zymomonas mobilis
ferredoxin
oxidoreductase
Improving ethanol Cloning and Heterologous expression of T. Hydrogen production decreased 1.5 [214]
yield transformation saccharolyticum ethanol production
operon (adhA, nfnAB, and adhEG544D)
Author Copy
122 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
al. [122] also analyzed the heterologous expression in C. thermocellum of the
genes xylA (xylose isomerase) and xylB (xylulokinase) from a thermophilic
anaerobic bacterium Thermoanaerobacter ethanolicus. The results from
this study suggested that the combined activity of these enzymes converted
xylose to xylulose-5-phosphate, which was then converted to ethanol
through the incorporation to PPP, glycolysis, and malate shunt pathway.
Searching for new metabolic reactions through enzyme engineering is one of
the expanding areas to improve actual scenario in the industrial sector [198].
Fossil fuels can potentially be replaced by biofuels in our daily lives, albeit
much work must develop to make the current biofuels production technology
economically feasible. Several considerations have to take in count for cost
reduction, such as maximum theorical yields in energy terms and cheaper
biomass feedstock fermentation. The transition from conventional fuels
to bioenergy systems requires the design and discovery of new metabolic
pathways, as well of new enzymes, new approaches in the bioprocess
engineering and development of microbial consortium capable of tolerate
high ethanol concentrations [223].
As described in sections above, recent advances and future perspectives
in microbial genetic engineering due to the fusion of biological approaches
derived from data of transcriptomics, proteomics, and metabolomics have
led to design novel synthetic circuits that might allow us to cross the gap
between the laboratory scale to the industrial marketplace [8].
Author Copy
an increase of ethanol production and yield by 40% and 22% respectively.
In addition, Xue et al. [224] used CRISPR/Cas9 technology to disrupt the
alcohol dehydrogenase 2 (ADH2) gene via complete deletion of the gen in S.
cerevisiae. Their results demonstrated an improvement up to 75% in ethanol
yield compared with the native strain.
5.6 CONCLUSIONS
KEYWORDS
• bioethanol
Apple Academic Press
• biofuels
• global transcription machinery engineering
• metabolic engineering
• prefoldin
• synthetic biology
Author Copy
REFERENCES
1. Shahbaz, M., Tiwari, A. K., & Nasir, M., (2013). The effects of financial development,
economic growth, coal consumption and trade openness on CO2 emissions in South
Africa. Energy Policy, 61, 1452–1459.
2. Awodumi, O. B., & Adewuyi, A. O., (2020). The role of non-renewable energy consump-
tion in economic growth and carbon emission: Evidence from oil producing economies
in Africa. Energy Strategy Reviews, 27, 100434.
3. Ching-Sung, T., Kwak, S., Turner, T. L., & Yong-Su, J., (2015). Yeast synthetic
biology toolbox and applications for biofuel production. FEMS Yeast Research,
15(1), 1–15.
4. Abdulkareem-Alsultan, G., Asikin-Mijan, N., Lee, H. V., & Taufiq-Yap, Y. H., (2020).
Biofuels: Past, present, future. In: Innovations in Sustainable Energy and Cleaner Environ-
ment (pp. 489–504). Springer, Singapore.
5. Vasconcelos, M. H., Mendes, F. M., Ramos, L., Dias, M. O. S., Bonomi, A., Jesus,
C. D. F., & Dos, S. J. C., (2020). Techno-economic assessment of bioenergy and
biofuel production in integrated sugarcane biorefinery: Identification of technological
bottlenecks and economic feasibility of dilute acid pretreatment. Energy, 117422.
6. Das, M., Patra, P., & Ghosh, A., (2020). Metabolic engineering for enhancing microbial
biosynthesis of advanced biofuels. Renewable and Sustainable Energy Reviews, 119,
109562.
7. Nielsen, J., & Keasling, J. D., (2016). Engineering cellular metabolism. Cell, 164(6),
1185–1197.
8. Madhavan, A., Jose, A. A., Binod, P., Sindhu, R., Sukumaran, R. K., Pandey, A., & Castro,
G. E., (2017). Synthetic biology and metabolic engineering approaches and its impact on
non-conventional yeast and biofuel production. Frontiers in Energy Research, 5, 8.
9. Adrio, J. L., & Demain, A. L., (2006). Genetic improvement of processes yielding
microbial products. FEMS Microbiology Reviews, 30(2), 187–214.
10. Lian, J., Mishra, S., & Zhao, H., (2018). Recent advances in metabolic engineering of
Saccharomyces cerevisiae: New tools and their applications. Metabolic Engineering,
50, 85–108.
11. Jiang, M., Stephanopoulos, G., & Pfeifer, B. A., (2012). Toward biosynthetic design
and implementation of Escherichia coli-derived paclitaxel and other heterologous
polyisoprene compounds. Applied and Environmental Microbiology, 78(8), 2497–2504.
12. Hwang, H. G., Kim, K. J., Lee, S. H., Kim, C. K., Min, C. K., Yun, J. M., & Son, Y.
Apple Academic Press
J., (2016). Recombinant glargine insulin production process using Escherichia coli. J
Microbiol. Biotechnol., 26(10), 1781–9.
13. Zhang, K., Lu, X., Li, Y., Jiang, X., Liu, L., & Wang, H., (2019). New technologies
provide more metabolic engineering strategies for bioethanol production in Zymomonas
mobilis. In: Applied Microbiology and Biotechnology (Vol. 103, No. 5, pp. 2087–2099).
https://doi.org/10.1007/s00253-019-09620-6.
14. Brown, S. D., Sander, K. B., Wu, C. W., & Guss, A. M., (2015). Clostridium thermocellum:
Engineered for the production of bioethanol. In: Direct Microbial Conversion of Biomass
to Advanced Biofuels (pp. 321–333). Elsevier.
Author Copy
15. Hopkins, M. M., Kraft, A., Martin, P. A., Nightingale, P., & Mahdi, S., (2007). Is the
Biotechnology Revolution a Myth? 1(17), 591–613.
16. Papagianni, M., (2017). Microbial bioprocesses. In: Current Developments in
Biotechnology and Bioengineering (pp. 45–72). Elsevier.
17. Doran, P. M., (2013). Bioprocess development: An interdisciplinary challenge.
Bioprocess Eng. Princ., 3–11.
18. Ko, Y. S., Kim, J. W., Lee, J. A., Han, T., Kim, G. B., Park, J. E., & Lee, S. Y., (2020).
Tools and strategies of systems metabolic engineering for the development of microbial
cell factories for chemical production. Chemical Society Reviews.
19. Singh, R., (2005). Hybrid membrane systems-applications and case studies. Hybrid
Membrane Systems for Water Purification, 131. Elsevier Science, Amsterdam.
20. Yang, S. T., (2007). Bioprocessing-from biotechnology to biorefinery. In: Bioprocessing
for Value-Added Products from Renewable Resources (pp. 1–24). Elsevier.
21. Stephanopoulos, G., Aristidou, A. A., & Nielsen, J., (1998). Metabolic Engineering:
Principles and Methodologies. Elsevier.
22. Vigneswaran, C., Ananthasubramanian, M., & Kandhavadivu, P., (2014). Enzymes
Technology in Bioprocessing of Textiles. WPI Publishing.
23. Ledesma-Amaro, R., Nikel, P. I., & Ceroni, F., (2020). Synthetic Biology-Guided
Metabolic Engineering. Frontiers in Bioengineering and Biotechnology, 8.
24. Qin, R., Zhong, C., Zong, G., Fu, J., Pang, X., & Cao, G., (2017). Improvement of
clavulanic acid production in Streptomyces clavuligerus F613-1 by using a claR-neo
reporter strategy. Electronic Journal of Biotechnology, 28, 41–46.
25. Parisutham, V., Kim, T. H., & Lee, S. K., (2014). Feasibilities of consolidated bioprocessing
microbes: From pretreatment to biofuel production. Bioresource Technology, 161, 431–440.
26. Biddy, M. J., Davis, R., Humbird, D., Tao, L., Dowe, N., Guarnieri, M. T., & Beckham,
G. T., (2016). The techno-economic basis for coproduct manufacturing to enable
hydrocarbon fuel production from lignocellulosic biomass. ACS Sustainable Chemistry
& Engineering, 4(6), 3196–3211.
27. Okamoto, K., Kondo-Okamoto, N., & Ohsumi, Y., (2009). Mitochondria-anchored
receptor Atg32 mediates degradation of mitochondria via selective autophagy.
Developmental Cell, 17(1), 87–97.
28. Eiyama, A., Kondo-Okamoto, N., & Okamoto, K., (2013). Mitochondrial degradation
during starvation is selective and temporally distinct from bulk autophagy in yeast.
FEBS Letters, 587(12), 1787–1792.
29. Shiroma, S., Jayakody, L. N., Horie, K., Okamoto, K., & Kitagaki, H., (2014). Enhancement
of ethanol fermentation in Saccharomyces cerevisiae sake yeast by disrupting mitophagy
function. Applied and Environmental Microbiology, 80(3), 1002–1012.
30. Peksel, A., Torres, N., Liu, J., Juneau, G., & Kubicek, C., (2002). 13 C-NMR analysis
Apple Academic Press
Author Copy
ethanol production from sugarcane molasses by industrially engineered Saccharomyces
cerevisiae via replacement of the PHO4 gene. RSC Advances, 10(4), 2267–2276.
34. Tan, M. J., Chen, X., Wang, Y. K., Liu, G. L., & Chi, Z. M., (2016a). Enhanced citric
acid production by a yeast yarrowia lipolytica over-expressing a pyruvate carboxylase
gene. Bioprocess and Biosystems Engineering, 39(8), 1289–1296.
35. Zhang, J., Wu, C., Wang, W., & Wei, D., (2018). Construction of enhanced transcriptional
activators for improving cellulase production in Trichoderma reesei RUT C30.
Bioresources and Bioprocessing, 5(1), 40.
36. Li, H., Zhang, Y., Gao, Y., Lan, Y., Yin, X., & Huang, L., (2016). Characterization of
UGPase from Aureobasidium pullulans NRRL Y-12974 and application in enhanced
pullulan production. Applied Biochemistry and Biotechnology, 178(6), 1141–1153.
37. Zhang, B., Lei, Z., Liu, Z. Q., & Zheng, Y. G., (2020a). Improvement of gibberellin produc-
tion by a newly isolated Fusarium fujikuroi mutant. Journal of Applied Microbiology.
38. Paudel, S., & Menze, M. A., (2014). Genetic engineering, a hope for sustainable biofuel
production. International Journal of Environment, 311.
39. Shanmugam, S., Ngo, H. H., & Wu, Y. R., (2019). Advanced CRISPR/Cas-based genome
editing tools for microbial biofuels production: A review. Renewable Energy. doi:
10.1016/j.renene.2019.10.107.
40. Dai, Z., Zhang, S., Yang, Q., Zhang, W., Qian, X., Dong, W., & Xin, F., (2018). Genetic
tool development and systemic regulation in biosynthetic technology. Biotechnology for
Biofuels, 11(1), 152.
41. Jagadevan, S., Banerjee, A., Banerjee, C., Guria, C., Tiwari, R., Baweja, M., & Shukla,
P., (2018). Recent developments in synthetic biology and metabolic engineering in
microalgae towards biofuel production. Biotechnology for Biofuels, 11(1), 185.
42. Javed, M. R., Noman, M., Shahid, M., Ahmed, T., Khurshid, M., Rashid, M. H., & Khan,
F., (2018). Current situation of biofuel production and its enhancement by CRISPR/
Cas9-mediated genome engineering of microbial cells. Microbiological Research. doi:
10.1016/j.micres.2018.10.010.
43. Dunlop, M. J., (2011). Engineering microbes for tolerance to next-generation biofuels.
Biotechnology for Biofuels, 4(1), 1–9.
44. Zingaro, K. A., & Terry, P. E., (2013). GroESL overexpression imparts Escherichia
coli tolerance to i-, n-, and 2-butanol, 1,2,4-butanetriol and ethanol with complex
and unpredictable patterns. Metabolic Engineering, 15, 196–205. doi: 10.1016/j.
ymben.2012.07.009.
45. Tomas, C. A., Welker, N. E., & Papoutsakis, E. T., (2003). Overexpression of groESL
in Clostridium acetobutylicum results in increased solvent production and tolerance,
prolonged metabolism, and changes in the cell’s transcriptional program. Applied and
Environmental Microbiology, 69(8), 4951–4965.
Apple Academic Press
46. Wang, S., He, Z., & Yuan, Q., (2017a). Xylose enhances furfural tolerance in candida
tropicalis by improving NADH recycle. Chemical Engineering Science, 158, 37–40.
47. Reyes, L. H., Abdelaal, A. S., & Kao, K. C., (2013). Genetic determinants for n-butanol
tolerance in evolved Escherichia coli mutants: Cross adaptation and antagonistic pleiotropy
between n-butanol and other stressors. Applied and Environmental Microbiology, 79(17),
5313–5320.
48. Mukhopadhyay, A., (2015). Tolerance engineering in bacteria for the production of
advanced biofuels and chemicals. Trends in Microbiology, 23(8), 498–508.
49. Tomas, C. A., Beamish, J., & Papoutsakis, E. T., (2004). Transcriptional analysis of
Author Copy
butanol stress and tolerance in Clostridium acetobutylicum. Journal of Bacteriology,
186(7), 2006–2018.
50. Brynildsen, M. P., & Liao, J. C., (2009). An integrated network approach identifies the
isobutanol response network of Escherichia coli. Molecular Systems Biology, 5(1), 277.
51. Rutherford, B. J., Dahl, R. H., Price, R. E., Szmidt, H. L., Benke, P. I., Mukhopadhyay,
A., & Keasling, J. D., (2010). Functional genomic study of exogenous n-butanol stress
in Escherichia coli. Applied and Environmental Microbiology, 76(6), 1935–1945.
52. Clark, D. S., Whitehead, T., Robb, F. T., Laksanalamai, P., & Jiemjit, A., (2014). U.S.
Patent No. 8,685,729. Washington, DC: U.S. Patent and Trademark Office.
53. Luan, G., Dong, H., Zhang, T., Lin, Z., Zhang, Y., Li, Y., & Cai, Z., (2014). Engineering
cellular robustness of microbes by introducing the GroESL chaperonins from
extremophilic bacteria. Journal of Biotechnology, 178, 38–40.
54. Abdelaal, A. S., Ageez, A. M., Abd El, A. E. H. A., & Abdallah, N. A., (2015). Genetic
improvement of n-butanol tolerance in Escherichia coli by heterologous overexpression
of groESL operon from Clostridium acetobutylicum. 3 Biotech, 5(4), 401–410.
55. Xu, G., Wu, A., Xiao, L., Han, R., & Ni, Y., (2019a). Enhancing butanol tolerance of
Escherichia coli reveals hydrophobic interaction of multi-tasking chaperone SecB.
Biotechnology for Biofuels, 12(1), 164.
56. Zhang, F., Qian, X., Si, H., Xu, G., Han, R., & Ni, Y., (2015). Significantly improved
solvent tolerance of Escherichia coli by global transcription machinery engineering.
Microbial Cell Factories, 14(1), 175.
57. Wang, M., Liu, L., Fan, L., & Tan, T., (2017b). CRISPRi based system for enhancing
1-butanol production in engineered Klebsiella pneumoniae. Process Biochemistry, 56,
139–146.
58. Otoupal, P. B., & Chatterjee, A., (2018). CRISPR gene perturbations provide insights for
improving bacterial biofuel tolerance. Frontiers in Bioengineering and Biotechnology,
6, 122.
59. Chen, S., Lesnik, E. A., Hall, T. A., Sampath, R., Griffey, R. H., Ecker, D. J., & Blyn,
L. B., (2002). A bioinformatics-based approach to discover small RNA genes in the
Escherichia coli genome. Biosystems, 65(2, 3), 157–177.
60. Erickson, K. E., Otoupal, P. B., & Chatterjee, A., (2017). Transcriptome-Level
Signatures in Gene Expression and Gene Expression Variability During Bacterial
Adaptive Evolution, 2(1). MSphere.
61. Huang, S., & Geng, A., (2020). High-copy genome integration of 2, 3-butanediol
biosynthesis pathway in Saccharomyces cerevisiae via in vivo DNA assembly and
replicative CRISPR-Cas9 mediated delta integration. Journal of Biotechnology, 310,
13–20.
Apple Academic Press
62. Kieboom, J., Dennis, J. J., De Bont, J. A., & Zylstra, G. J., (1998). Identification and
molecular characterization of an efflux pump involved in Pseudomonas putida S12
solvent tolerance. Journal of Biological Chemistry, 273(1), 85–91.
63. Segura, A., Duque, E., Mosqueda, G., Ramos, J. L., & Junker, F., (1999). Multiple
responses of gram-negative bacteria to organic solvents. Environmental Microbiology,
1(3), 191–198.
64. Lee, J. Y., Geraldi, A., Rahman, Z., Lee, J. H., & Kim, S. C., (2015). Improved n-butanol
tolerance in Escherichia coli by controlling membrane related functions. Journal of
Biotechnology, 204, 33–44.
Author Copy
65. Jiménez-Bonilla, P., Zhang, J., Wang, Y., Blersch, D., de-Bashan, L. E., Guo, L., &
Wang, Y., (2020). Enhancing the tolerance of Clostridium saccharoperbutylacetonicum
to lignocellulosic-biomass-derived inhibitors for efficient biobutanol production by
overexpressing efflux pumps genes from Pseudomonas putida. Bioresource Technology,
123532.
66. Ling, H., Chen, B., Kang, A., Lee, J. M., & Chang, M. W., (2013). Transcriptome
response to alkane biofuels in Saccharomyces cerevisiae: Identification of efflux pumps
involved in alkane tolerance. Biotechnology for Biofuels, 6(1), 95.
67. Boyarskiy, S., Davis, L. S., Kong, N., & Tullman-Ercek, D., (2016). Transcriptional
feedback regulation of efflux protein expression for increased tolerance to and production
of n -butanol. Metabolic Engineering, 33, 130–137. doi: 10.1016/j.ymben.2015.11.005.
68. Fisher, M. A., Boyarskiy, S., Yamada, M. R., Kong, N., Bauer, S., & Tullman-Ercek, D.,
(2014). Enhancing tolerance to short-chain alcohols by engineering the Escherichia coli
AcrB efflux pump to secrete the non-native substrate n-butanol. ACS Synthetic Biology,
3(1), 30–40.
69. Foo, J. L., & Leong, S. S. J., (2013). Directed evolution of an E. coli inner membrane
transporter for improved efflux of biofuel molecules. Biotechnology for Biofuels, 6(1),
1–12.
70. He, X., Xue, T., Ma, Y., Zhang, J., Wang, Z., Hong, J., Hui, L., et al., (2019). Identification
of functional butanol-tolerant genes from Escherichia coli mutants derived from error-
prone PCR-based whole-genome shuffling. Biotechnology for Biofuels, 12(1), 73.
71. Bhagat, R., Panakkal, H., Gupta, I., & Ingle, A. P., (2020). Recent advances in the
production of biodiesel using lignocellulosic biomass. Lignocellulosic Biorefining
Technologies, 69–85. doi: 10.1002/9781119568858.ch5.
72. Palmqvist, E., & Hahn-Hägerdal, B., (2000). Fermentation of lignocellulosic
hydrolysates. II: Inhibitors and mechanisms of inhibition. Bioresource Technology,
74(1), 25–33. doi: 10.1016/s0960-8524(99)00161-3.
73. Almeida, J. R., Modig, T., Petersson, A., Hähn-Hägerdal, B., Lidén, G., & Gorwa-
Grauslund, M. F., (2007). Increased tolerance and conversion of inhibitors in lignocel-
lulosic hydrolysates by Saccharomyces cerevisiae. Journal of Chemical Technology &
Biotechnology: International Research in Process, Environmental & Clean Technology,
82(4), 340–349.
74. Gorsich, S. W., Dien, B. S., Nichols, N. N., Slininger, P. J., Liu, Z. L., & Skory, C.
D., (2006). Tolerance to furfural-induced stress is associated with pentose phosphate
pathway genes ZWF1, GND1, RPE1, and TKL1 in Saccharomyces cerevisiae. Applied
Microbiology and Biotechnology, 71(3), 339–349.
75. Miller, E. N., Jarboe, L. R., Turner, P. C., Pharkya, P., Yomano, L. P., York, S. W.,
& Ingram, L. O., (2009). Furfural inhibits growth by limiting sulfur assimilation in
Apple Academic Press
Author Copy
factors on ethanol yields from lignocellulosic biomass by Thermoanaerobacterium
AK17. Biotechnology and Bioengineering, 109(3), 686–694.
79. Seo, H. M., Jeon, J. M., Lee, J. H., Song, H. S., Joo, H. B., Park, S. H., & Lee, H., (2016).
Combinatorial application of two aldehyde oxidoreductases on isobutanol production in
the presence of furfural. Journal of Industrial Microbiology & Biotechnology, 43(1),
37–44.
80. Song, H. S., Jeon, J. M., Kim, H. J., Bhatia, S. K., Sathiyanarayanan, G., Kim, J., &
Yang, Y. H., (2017). Increase in furfural tolerance by combinatorial overexpression of
NAD salvage pathway enzymes in engineered isobutanol-producing E. coli. Bioresource
Technology, 245, 1430–1435.
81. Oh, E. J., Wei, N., Kwak, S., Kim, H., & Jin, Y. S., (2019). Overexpression of RCK1
improves acetic acid tolerance in Saccharomyces cerevisiae. Journal of Biotechnology.
doi: 10.1016/j.jbiotec.2018.12.013.
82. Liu, C. G., Cao, L. Y., Wen, Y., Li, K., Mehmood, M. A., Zhao, X. Q., & Bai, F. W.,
(2020a). Intracellular redox manipulation of Zymomonas mobilis for improving toler-
ance against lignocellulose hydrolysate-derived stress. Chemical Engineering Science,
115933.
83. Ujor, V., Agu, C. V., Gopalan, V., & Ezeji, T. C., (2014). Glycerol supplementation of
the growth medium enhances in situ detoxification of furfural by Clostridium beijer-
inckii during butanol fermentation. Applied Microbiology and Biotechnology, 98(14),
6511–6521.
84. Agu, C. V., Ujor, V., & Ezeji, T. C., (2019). Metabolic engineering of Clostridium
beijerinckii to improve glycerol metabolism and furfural tolerance. Biotechnology for
Biofuels, 12(1), 50.
85. Caspeta, L., Chen, Y., Ghiaci, P., Feizi, A., Buskov, S., Hallström, B. M., & Nielsen, J.,
(2014). Altered sterol composition renders yeast thermotolerant. Science, 346(6205),
75–78.
86. Li, P., Fu, X., Zhang, L., Zhang, Z., Li, J., & Li, S., (2017). The transcription factors
Hsf1 and Msn2 of thermotolerant Kluyveromyces marxianus promote cell growth and
ethanol fermentation of Saccharomyces cerevisiae at high temperatures. Biotechnology
for Biofuels, 10(1), 1–13.
87. Li, P., Fu, X., Zhang, L., & Li, S., (2019). CRISPR/Cas-based screening of a gene
activation library in Saccharomyces cerevisiae identifies a crucial role of OLE 1 in
thermotolerance. Microbial. Biotechnology, 12(6), 1154–1163.
88. Tessarz, P., Schwarz, M., Mogk, A., & Bukau, B., (2009). The yeast AAA+ chaperone
Hsp104 is part of a network that links the actin cytoskeleton with the inheritance of
damaged proteins. Molecular and Cellular Biology, 29(13), 3738–3745.
89. De Virgilio, C., Hottiger, T., Dominguez, J., Boller, T., & Wiemken, A., (1994). The
Apple Academic Press
Author Copy
92. Benjaphokee, S., Koedrith, P., Auesukaree, C., Asvarak, T., Sugiyama, M., Kaneko,
Y., & Harashima, S., (2012a). CDC19 encoding pyruvate kinase is important for high-
temperature tolerance in Saccharomyces cerevisiae. New Biotechnology, 29(2), 166–176.
93. Nasution, O., Lee, J., Srinivasa, K., Choi, I. G., Lee, Y. M., Kim, E., & Kim, W., (2015).
Loss of D fg5 glycosylphosphatidylinositol-anchored membrane protein confers
enhanced heat tolerance in Saccharomyces cerevisiae. Environmental Microbiology,
17(8), 2721–2734.
94. Jia, H., Sun, X., Sun, H., Li, C., Wang, Y., Feng, X., & Li, C., (2016). Intelligent microbial
heat-regulating engine (IMHeRE) for improved thermo-robustness and efficiency of
bioconversion. ACS Synthetic Biology, 5(4), 312–320. doi: 10.1021/acssynbio.5b00158.
95. Xu, K., Gao, L., Hassan, J. U., Zhao, Z., Li, C., Huo, Y. X., & Liu, G., (2018). Improving
the thermo-tolerance of yeast base on the antioxidant defense system. Chemical
Engineering Science, 175, 335–342. doi: 10.1016/j.ces.2017.10.016.
96. Li, K., Zhang, J. W., Liu, C. G., Mehmood, M. A., & Bai, F. W., (2020). Elucidating the
molecular mechanism of TEMPOL-mediated improvement on tolerance under oxidative
stress in Saccharomyces cerevisiae. Chemical Engineering Science, 211, 115306.
97. Tao, X., Xu, T., Kempher, M. L., Liu, J., & Zhou, J., (2020). Precise promoter integration
improves cellulose bioconversion and thermotolerance in Clostridium cellulolyticum.
Metabolic Engineering.
98. Zhao, X., Zhang, L., & Liu, D., (2012). Biomass recalcitrance. Part I: The chemical
compositions and physical structures affecting the enzymatic hydrolysis of lignocellulose.
Biofuels, Bioproducts and Biorefining, 6(4), 465–482.
99. Ballesteros, L. F., Michelin, M., Vicente, A. A., Teixeira, J. A., & Cerqueira, M. Â., (2018).
Lignocellulosic materials: Sources and processing technologies. In: Lignocellulosic
Materials and Their Use in Bio-based Packaging (pp. 13–33). Springer, Cham.
100. Bhatia, S. K., Jagtap, S. S., Bedekar, A. A., Bhatia, R. K., Patel, A. K., Pant, D., & Yang,
Y. H., (2020). Recent developments in pretreatment technologies on lignocellulosic
biomass: Effect of key parameters, technological improvements, and challenges.
Bioresource Technology, 300, 122724.
101. Xu, N., Liu, S., Xin, F., Jia, H., Xu, J., Jiang, M., & Dong, W., (2019b). Biomethane
production from lignocellulose: Biomass recalcitrance and its impacts on anaerobic
digestion. Frontiers in Bioengineering and Biotechnology, 7, 191.
102. Lu, S., Li, L., & Zhou, G., (2010). Genetic modification of wood quality for second-
generation biofuel production. GM Crops, 1(4), 230–236.
103. Sharma, R., Joshi, R., & Kumar, D., (2020a). Present status and future prospect
of genetic and metabolic engineering for biofuels production from lignocellulosic
biomass. In: Genetic and Metabolic Engineering for Improved Biofuel Production from
Lignocellulosic Biomass (pp. 171–192). Elsevier.
Apple Academic Press
104. Franke, R., McMichael, C. M., Meyer, K., Shirley, A. M., Cusumano, J. C., & Chapple,
C., (2000). Modified lignin in tobacco and poplar plants over-expressing the Arabidopsis
gene encoding ferulate 5-hydroxylase. The Plant Journal, 22(3), 223–234.
105. Abdulrazzak, N., Pollet, B., Ehlting, J., Larsen, K., Asnaghi, C., Ronseau, S., & Ullmann,
P., (2006). A coumaroyl-ester-3-hydroxylase insertion mutant reveals the existence of
nonredundant meta-hydroxylation pathways and essential roles for phenolic precursors
in cell expansion and plant growth. Plant Physiology, 140(1), 30–48.
106. Ralph, J., (2006). What makes a good monolignol substitute. The Science and Lore of
the Plant Cell Wall Biosynthesis, Structure and Function, 285–293.
Author Copy
107. Besseau, S., Hoffmann, L., Geoffroy, P., Lapierre, C., Pollet, B., & Legrand, M., (2007).
Flavonoid accumulation in Arabidopsis repressed in lignin synthesis affects auxin
transport and plant growth. The Plant Cell, 19(1), 148–162.
108. Coleman, H. D., Park, J. Y., Nair, R., Chapple, C., & Mansfield, S. D., (2008). RNAi-
mediated suppression of p-coumaroyl-CoA 3′-hydroxylase in hybrid poplar impacts
lignin deposition and soluble secondary metabolism. Proceedings of the National
Academy of Sciences, 105(11), 4501–4506.
109. Chen, F., & Dixon, R. A., (2007). Lignin modification improves fermentable sugar
yields for biofuel production. Nature Biotechnology, 25(7), 759–761.
110. Shafrin, F., Ferdous, A. S., Sarkar, S. K., Ahmed, R., Hossain, K., Sarker, M., & Khan,
H., (2017). Modification of monolignol biosynthetic pathway in jute: Different gene,
different consequence. Scientific Reports, 7(1), 1–12.
111. Jung, J. H., Fouad, W. M., Vermerris, W., Gallo, M., & Altpeter, F., (2012). RNAi
suppression of lignin biosynthesis in sugarcane reduces recalcitrance for biofuel production
from lignocellulosic biomass. Plant Biotechnology Journal, 10(9), 1067–1076.
112. Ragauskas, A. J., Williams, C. K., Davison, B. H., Britovsek, G., Cairney, J., Eckert, C.
A., & Mielenz, J. R., (2006). The path forward for biofuels and biomaterials. Science,
311(5760), 484–489.
113. Studer, M. H., DeMartini, J. D., Davis, M. F., Sykes, R. W., Davison, B., Keller, M., &
Wyman, C. E., (2011). Lignin content in natural Populus variants affects sugar release.
Proceedings of the National Academy of Sciences, 108(15), 6300–6305.
114. Zhang, J., Xie, M., Li, M., Ding, J., Pu, Y., Bryan, A. C., & Lindquist, E. A., (2020b).
Overexpression of a prefoldin β subunit gene reduces biomass recalcitrance in the
bioenergy crop Populus. Plant Biotechnology Journal, 18(3), 859–871.
115. Sakamoto, S., Kamimura, N., Tokue, Y., Nakata, M. T., Yamamoto, M., Hu, S., & Kajita,
S., (2020). Identification of enzymatic genes with the potential to reduce biomass
recalcitrance through lignin manipulation in Arabidopsis. Biotechnology for Biofuels,
13, 1–16.
116. Tesfaw, A., & Assefa, F., (2014). Current trends in bioethanol production by Saccharomyces
cerevisiae: Substrate, inhibitor reduction, growth variables, coculture, and immobilization.
International Scholarly Research Notices, 2014.
117. Azhar, S. H. M., Abdulla, R., Jambo, S. A., Marbawi, H., Gansau, J. A., Faik, A. A.
M., & Rodrigues, K. F., (2017). Yeasts in sustainable bioethanol production: A review.
Biochemistry and Biophysics Reports, 10, 52–61.
118. Deparis, Q., Claes, A., Foulquié-Moreno, M. R., & Thevelein, J. M., (2017). Engineering
tolerance to industrially relevant stress factors in yeast cell factories. FEMS Yeast
Research, 17(4).
119. Lau, M. W., Gunawan, C., Balan, V., & Dale, B. E., (2010). Comparing the fermentation
Apple Academic Press
Author Copy
122. Banerjee, S., Mishra, G., & Roy, A., (2019). Metabolic engineering of bacteria for
renewable bioethanol production from cellulosic biomass. Biotechnology and Bioprocess
Engineering, 1–21.
123. Banat, I. M., Nigam, P., Singh, D., Marchant, R., & McHale, A. P., (1998). Ethanol
production at elevated temperatures and alcohol concentrations: Part I–yeasts in general.
world Journal of Microbiology and Biotechnology, 14(6), 809–821.
124. Yamaoka, C., Kurita, O., & Kubo, T., (2014). Improved ethanol tolerance of Saccharo-
myces cerevisiae in mixed cultures with Kluyveromyces lactis on high-sugar fermentation.
Microbiological Research, 169(12), 907–914.
125. Ansanay-Galeote, V., Blondin, B., Dequin, S., & Sablayrolles, J. M., (2001). Stress
effect of ethanol on fermentation kinetics by stationary-phase cells of Saccharomyces
cerevisiae. Biotechnology Letters, 23(9), 677–681.
126. Auesukaree, C., (2017). Molecular mechanisms of the yeast adaptive response and
tolerance to stresses encountered during ethanol fermentation. Journal of Bioscience
and Bioengineering, 124(2), 133–142.
127. Riles, L., & Fay, J. C., (2019). Genetic basis of variation in heat and ethanol tolerance in
Saccharomyces cerevisiae. G3: Genes, Genomes, Genetics, 9(1), 179–188.
128. Hu, X. H., Wang, M. H., Tan, T., Li, J. R., Yang, H., Leach, L., & Luo, Z. W., (2007).
Genetic dissection of ethanol tolerance in the budding yeast Saccharomyces cerevisiae.
Genetics, 175(3), 1479–1487.
129. Stanley, D., Bandara, A., Fraser, S., Chambers, P. J., & Stanley, G. A., (2010). The
ethanol stress response and ethanol tolerance of Saccharomyces cerevisiae. J. Appl.
Microbiol., 109(1), 13–24.
130. Charoenbhakdi, S., Dokpikul, T., Burphan, T., Techo, T., & Auesukaree, C., (2016).
Vacuolar H+-ATPase protects Saccharomyces cerevisiae cells against ethanol-induced
oxidative and cell wall stresses. Applied and Environmental Microbiology, 82(10),
3121–3130.
131. Pascual, C., Alonso, A., Garcia, I., Romay, C., & Kotyk, A., (1988). Effect of ethanol
on glucose transport, key glycolytic enzymes, and proton extrusion in Saccharomyces
cerevisiae. Biotechnology and Bioengineering, 32(3), 374–378.
132. Ma, M., & Liu, Z. L., (2010). Mechanisms of ethanol tolerance in Saccharomyces
cerevisiae. Applied Microbiology and Biotechnology, 87(3), 829–845.
133. Hirasawa, T., Yoshikawa, K., Nakakura, Y., Nagahisa, K., Furusawa, C., Katakura, Y.,
& Shioya, S., (2007). Identification of target genes conferring ethanol stress tolerance
135. Nuwamanya, E., Chiwona-Karltun, L., Kawuki, R. S., & Baguma, Y., (2012). Bio-ethanol
production from non-food parts of cassava (Manihot esculenta Crantz). Ambio, 41(3),
262–270.
136. Benjaphokee, S., Hasegawa, D., Yokota, D., Asvarak, T., Auesukaree, C., Sugiyama, M.,
& Harashima, S., (2012). Highly efficient bioethanol production by a Saccharomyces
cerevisiae strain with multiple stress tolerance to high temperature, acid and ethanol.
New Biotechnology, 29(3), 379–386.
137. Shahsavarani, H., Sugiyama, M., Kaneko, Y., Chuenchit, B., & Harashima, S., (2012).
Superior thermotolerance of Saccharomyces cerevisiae for efficient bioethanol
Author Copy
fermentation can be achieved by overexpression of RSP5 ubiquitin ligase. Biotechnology
Advances, 30(6), 1289–1300.
138. Mahmud, S. A., Hirasawa, T., & Shimizu, H., (2010). Differential importance of trehalose
accumulation in Saccharomyces cerevisiae in response to various environmental stresses.
Journal of Bioscience and Bioengineering, 109(3), 262–266.
139. Wiemken, A., (1990). Trehalose in yeast, stress protectant rather than reserve carbohydrate.
Antonie van Leeuwenhoek, 58(3), 209–217.
140. Singer, M. A., & Lindquist, S., (1998). Thermotolerance in Saccharomyces cerevisiae:
The yin and yang of trehalose. Trends in Biotechnology, 16(11), 460–468.
141. Vianna, C. R., Silva, C. L., Neves, M. J., & Rosa, C. A., (2008). Saccharomyces cerevisiae
strains from traditional fermentations of Brazilian cachaca: Trehalose metabolism, heat
and ethanol resistance. Antonie Van Leeuwenhoek, 93(1, 2), 205–217.
142. Cray, J. A., Stevenson, A., Ball, P., Bankar, S. B., Eleutherio, E. C., Ezeji, T. C., &
Hallsworth, J. E., (2015). Chaotropicity: A key factor in product tolerance of biofuel-
producing microorganisms. Current Opinion in Biotechnology, 33, 228–259.
143. Nwaka, S., Mechler, B., & Holzer, H., (1996). Deletion of the ATH1 gene in Saccharomyces
cerevisiae prevents growth on trehalose. FEBS Letters, 386(2, 3), 235–238.
144. Jung, Y. J., & Park, H. D., (2005). Antisense-mediated inhibition of acid trehalase
(ATH1) gene expression promotes ethanol fermentation and tolerance in Saccharomyces
cerevisiae. Biotechnology Letters, 27(23, 24), 1855–1859.
145. Caspeta, L., Buijs, N. A., & Nielsen, J., (2013). The role of biofuels in the future energy
supply. Energy & Environmental Science, 6(4), 1077–1082.
146. Caspeta, L., Castillo, T., & Nielsen, J., (2015). Modifying yeast tolerance to inhibitory
conditions of ethanol production processes. Frontiers in Bioengineering and Biotech-
nology, 3, 184.
147. Kumar, A., Singh, L. K., & Ghosh, S., (2009). Bioconversion of lignocellulosic fraction
of water-hyacinth (Eichhornia crassipes) hemicellulose acid hydrolysate to ethanol by
Pichia stipitis. Bioresource technology, 100(13), 3293–3297.
148. Laluce, C., Schenberg, A. C. G., Gallardo, J. C. M., Coradello, L. F. C., & Pombeiro-
Sponchiado, S. R., (2012). Advances and developments in strategies to improve strains
of Saccharomyces cerevisiae and processes to obtain the lignocellulosic ethanol: A
review. Applied Biochemistry and Biotechnology, 166(8), 1908–1926.
149. Komesu, A., Oliveira, J., Neto, J. M., Penteado, E. D., Diniz, A. A. R., & Da Silva, M.
L. H., (2020). Xylose fermentation to bioethanol production using genetic engineering
microorganisms. In: Genetic and Metabolic Engineering for Improved Biofuel Production
from Lignocellulosic Biomass (pp. 143–154). Elsevier.
150. Mussatto, S. I., Machado, E. M., Carneiro, L. M., & Teixeira, J. A., (2012). Sugars
metabolism and ethanol production by different yeast strains from coffee industry
Apple Academic Press
Author Copy
xylose in engineered Saccharomyces cerevisiae strains: Current state and perspectives.
Applied Microbiology and Biotechnology, 84(1), 37–53.
154. Hector, R. E., Dien, B. S., Cotta, M. A., & Mertens, J. A., (2013). Growth and fermenta-
tion of D-xylose by Saccharomyces cerevisiae expressing a novel D-xylose isomerase
originating from the bacterium Prevotella ruminicola TC2-24. Biotechnology for
Biofuels, 6(1), 1–12.
155. Katahira, S., Mizuike, A., Fukuda, H., & Kondo, A., (2006). Ethanol fermentation
from lignocellulosic hydrolysate by a recombinant xylose-and cellooligosaccharide-
assimilating yeast strain. Applied Microbiology and Biotechnology, 72(6), 1136–1143.
156. Karhumaa, K., Hahn-Hägerdal, B., & Gorwa-Grauslund, M. F., (2005). Investigation
of limiting metabolic steps in the utilization of xylose by recombinant Saccharomyces
cerevisiae using metabolic engineering. Yeast, 22(5), 359–368.
157. Fang, T., Yan, H., Li, G., Chen, W., Liu, J., & Jiang, L., (2020). Chromatin remodeling
complexes are involvesd in the regulation of ethanol production during static fermentation
in budding yeast. Genomics, 112(2), 1674–1679.
158. Fukuda, A., Kuriya, Y., Konishi, J., Mutaguchi, K., Uemura, T., Miura, D., & Okamoto,
M., (2019). Kinetic modeling and sensitivity analysis for higher ethanol production
in self-cloning xylose-using Saccharomyces cerevisiae. Journal of Bioscience and
Bioengineering, 127(5), 563–569.
159. Xiong, M., Chen, G., & Barford, J., (2014). Genetic engineering of yeasts to improve
ethanol production from xylose. Journal of the Taiwan Institute of Chemical Engineers,
45(1), 32–39.
160. Liu, T., Huang, S., & Geng, A., (2018). Recombinant diploid Saccharomyces cerevisiae
strain development for rapid glucose and xylose co-fermentation. Fermentation, 4(3), 59.
161. Yang, P., Zhang, H., & Jiang, S., (2016). Construction of recombinant sestc
Saccharomyces cerevisiae for consolidated bioprocessing, cellulase characterization,
and ethanol production by in situ fermentation. 3 Biotech, 6(2), 192.
162. Ko, J. K., Jung, J. H., Altpeter, F., Kannan, B., Kim, H. E., Kim, K. H., & Lee, S. M., (2018).
Largely enhanced bioethanol production through the combined use of lignin-modified
sugarcane and xylose fermenting yeast strain. Bioresource Technology, 256, 312–320.
163. Qin, L., Dong, S., Yu, J., Ning, X., Xu, K., Zhang, S. J., & Li, C., (2020). Stress-driven
dynamic regulation of multiple tolerance genes improves robustness and productive
capacity of Saccharomyces cerevisiae in industrial lignocellulose fermentation. Metabolic
Engineering.
164. Ohta, E., Nakayama, Y., Mukai, Y., Bamba, T., & Fukusaki, E., (2016). Metabolomic
approach for improving ethanol stress tolerance in Saccharomyces cerevisiae. Journal
of Bioscience and Bioengineering, 121(4), 399–405.
165. Sun, Y., Xue, Q., Hou, J., Kong, M., Li, X., He, B., & Cao, L., (2020). Research Square.
Apple Academic Press
Author Copy
168. Huang, S., Xue, T., Wang, Z., Ma, Y., He, X., Hong, J., Zou, S., Song, H., & Zhang,
M., (2018). Furfural-tolerant Zymomonas mobilis derived from error-prone PCR-based
whole genome shuffling and their tolerant mechanism. Applied Microbiology and
Biotechnology, 102(7), 3337–3347.
169. Fuchino, K., Kalnenieks, U., Rutkis, R., Grube, M., & Bruheim, P., (2020). Metabolic
profiling of glucose-fed metabolically active resting Zymomonas mobilis strains.
Metabolites, 10(3), 6–8. https://doi.org/10.3390/metabo10030081.
170. Liu, Y., Ghosh, I. N., Martien, J., Zhang, Y., Amador-Noguez, D., & Landick, R.,
(2020b). Regulated redirection of central carbon flux enhances anaerobic production of
bioproducts in Zymomonas mobilis. Metabolic Engineering, 61, 261–274. https://doi.
org/10.1016/j.ymben.2020.06.005.
171. Sarkar, S., Mukherjee, A., Das, S., Ghosh, B., Chaudhuri, S., Bhattacharya, D., Sarbajna,
A., & Gachhui, R., (2019). Nitrogen deprivation elicits dimorphism, capsule biosynthesis
and autophagy in Papiliotrema laurentii strain RY1. Micron, 124, 102708. https://doi.
org/10.1016/j.micron.2019.102708.
172. Ma, K., Ruan, Z., Shui, Z., Wang, Y., Hu, G., & He, M., (2016). Open fermentative
production of fuel ethanol from food waste by an acid-tolerant mutant strain of
Zymomonas mobilis. Bioresource Technology, 203, 295–302. https://doi.org/10.1016/j.
biortech.2015.12.054.
173. Todhanakasem, T., Wu, B., & Simeon, S., (2020). Perspectives and new directions for
bioprocess optimization using Zymomonas mobilis in the ethanol production. World Journal
of Microbiology and Biotechnology, 36(8), 1–16. https://doi.org/10.1007/s11274-020-
02885-4.
174. Yin, H., Zhang, R., Xia, M., Bai, X., Mou, J., Zheng, Y., & Wang, M., (2017). Effect of
aspartic acid and glutamate on metabolism and acid stress resistance of Acetobacter pasteur-
ianus. Microbial Cell Factories, 16(1), 1–14. https://doi.org/10.1186/s12934-017-0717-6.
175. Guan, N., & Liu, L., (2020). Microbial response to acid stress: Mechanisms and
applications. Applied Microbiology and Biotechnology, 104(1), 51–65. -https://doi.org/
10.1007/s00253-019-10226–1.
176. Todhanakasem, T., Yodsanga, S., Sowatad, A., Kanokratana, P., Thanonkeo, P., &
Champreda, V., (2018). Inhibition analysis of inhibitors derived from lignocellulose
pretreatment on the metabolic activity of Zymomonas mobilis biofilm and planktonic
cells and the proteomic responses. Biotechnology and Bioengineering, 115(1), 70–81.
https://doi.org/10.1002/bit.26449.
177. Liu, Y. F., Hsieh, C. W., Chang, Y. S., & Wung, B. S., (2017). Effect of acetic acid on
ethanol production by Zymomonas mobilis mutant strains through continuous adaptation.
BMC Biotechnology, 17(1), 1–10. https://doi.org/10.1186/s12896-017-0385-y.
178. Wu, B., Qin, H., Yang, Y., Duan, G., Yang, S., Xin, F., Zhao, C., et al., (2019). Engineered
Apple Academic Press
Zymomonas mobilis tolerant to acetic acid and low pH via multiplex atmospheric and
room temperature plasma mutagenesis. Biotechnology for Biofuels, 12(1), 1–13. https://
doi.org/10.1186/s13068-018-1348-9.
179. Singh, B., Kumar, P., Yadav, A., & Datta, S., (2019). Degradation of fermentation inhibi-
tors from lignocellulosic hydrolysate liquor using immobilized bacterium, Bordetella
sp. BTIITR. Chemical Engineering Journal, 361, 1152–1160. https://doi.org/10.1016/j.
cej.2018.12.168.
180. He, M. X., Wu, B., Qin, H., Ruan, Z. Y., Tan, F. R., Wang, J. L., Shui, Z. X., et al., (2014).
Zymomonas mobilis: A novel platform for future biorefineries. In: Biotechnology for
Author Copy
Biofuels (Vol. 7, No. 1, pp. 1–15).
181. Kurgan, G., Panyon, L. A., Rodriguez-Sanchez, Y., Pacheco, E., Nieves, L. M., Mann,
R., Nielsen, D. R., & Wanga, X., (2019). Bioprospecting of native efflux pumps to
enhance furfural tolerance in ethanologenic Escherichia coli. Applied and Environmental
Microbiology, 85(6), 1–11. https://doi.org/10.1128/AEM.02985-18.
182. Wang, W., Wu, B., Qin, H., Liu, P., Qin, Y., Duan, G., Hu, G., & He, M., (2019).
Genome shuffling enhances stress tolerance of Zymomonas mobilis to two inhibitors.
Biotechnology for Biofuels, 12(1), 1–12. https://doi.org/10.1186/s13068-019-1631-4.
183. Cao, H., Wei, D., Yang, Y., Shang, Y., Li, G., Zhou, Y., Ma, Q., & Xu, Y., (2017). Systems-
level understanding of ethanol-induced stresses and adaptation in E. coli. Scientific
Reports, pp. 1–15. https://doi.org/10.1038/srep44150.
184. Skupin, P., & Metzger, M., (2017). Stability analysis of the continuous ethanol
fermentation process with a delayed product inhibition. Applied Mathematical
Modelling, 49, 48–58. https://doi.org/10.1016/j.apm.2017.04.025.
185. Carreón-Rodríguez, O. E., Gutiérrez-Ríos, R. M., Acosta, J. L., Martinez, A., & Cevallos,
M. A., (2019). Phenotypic and genomic analysis of Zymomonas mobilis ZM4 mutants
with enhanced ethanol tolerance. Biotechnology Reports, 23. https://doi.org/10.1016/j.
btre.2019.e00328.
186. Tan, F., Wu, B., Dai, L., Qin, H., Shui, Z., Wang, J., Zhu, Q., Hu, G., & He, M., (2016b).
Using global transcription machinery engineering (gTME) to improve ethanol tolerance
of Zymomonas mobilis. Microbial Cell Factories, 15(1), 1–9. https://doi.org/10.1186/
s12934-015-0398-y.
187. Kim, D., (2018). Physico-chemical conversion of lignocellulose: Inhibitor effects and
detoxification strategies: A mini review. Molecules, 23(2). https://doi.org/10.3390/
molecules23020309.
188. Qiu, Z., Fang, C., Gao, Q., & Bao, J., (2020). A short-chain dehydrogenase plays a key
role in cellulosic D-lactic acid fermentability of Pediococcus acidilactici. Bioresource
Technology, 297, 122473. https://doi.org/10.1016/j.biortech.2019.122473.
189. Yi, X., Gao, Q., & Bao, J., (2019). Expressing an oxidative dehydrogenase gene in
ethanologenic strain Zymomonas mobilis promotes the cellulosic ethanol fermentability.
Journal of Biotechnology, 303, 1–7. https://doi.org/10.1016/j.jbiotec.2019.07.005.
190. Kopp, D., & Sunna, A. (2020). Alternative carbohydrate pathways–enzymes, functions
and engineering. Critical Reviews in Biotechnology, 40(7), 895–912.
191. Mazaheri, D., & Pirouzi, A., (2020). Valorization of Zymomonas mobilis for bioethanol
production from potato peel: Fermentation process optimization. Biomass Conversion
and Biorefinery.
192. Palamae, S., Choorit, W., Chatsungnoen, T., & Chisti, Y., (2020). Simultaneous nitrogen
Apple Academic Press
Author Copy
195. Duan, G., Wu, B., Qin, H., Wang, W., Tan, Q., Dai, Y., Qin, Y., Tan, F., Hu, G., & He, M.,
(2019). Replacing water and nutrients for ethanol production by ARTP derived biogas
slurry tolerant Zymomonas mobilis strain. Biotechnology for Biofuels, 12(1), 1–12.
196. Robak, K., & Balcerek, M., (2018). Review of second-generation bioethanol production
from residual biomass. Food technology and Biotechnology, 56(2), 174–187.
197. Valta, K., Papadaskalopoulou, C., Dimarogona, M., & Topakas, E., (2019). Bioethanol from
waste-prospects and challenges of current and emerging technologies. Byproducts from
Agriculture and Fisheries: Adding Value for Food, Feed, Pharma, and Fuels, 421–456.
198. Choi, K. R., Jiao, S., & Lee, S. Y., (2020). Metabolic engineering strategies toward
production of biofuels. Current Opinion in Chemical Biology, 59, 1–14.
199. Sharma, B., Larroche, C., & Dussap, C. G., (2020b). Comprehensive assessment of 2G
bioethanol production. Bioresource Technology, 123630.
200. Di Donato, P., Finore, I., Poli, A., Nicolaus, B., & Lama, L., (2019). The production
of second-generation bioethanol: The biotechnology potential of thermophilic bacteria.
Journal of Cleaner Production, 233, 1410–1417.
201. Lynd, L. R., Van, Z. W. H., McBride, J. E., & Laser, M., (2005). Consolidated bioprocessing
of cellulosic biomass: An update. Current Opinion in Biotechnology, 16(5), 577–583.
202. Poudel, S., Giannone, R. J., Rodriguez, M., Raman, B., Martin, M. Z., Engle, N. L., &
Ussery, D., (2017). Integrated omics analyses reveal the details of metabolic adaptation
of Clostridium thermocellum to lignocellulose-derived growth inhibitors released during
the deconstruction of switchgrass. Biotechnology for Biofuels, 10(1), 14.
203. Cui, J., Stevenson, D., Korosh, T., Amador-Noguez, D., Olson, D. G., & Lynd, L. R., (2020).
Developing a cell-free extract reaction (CFER) system in Clostridium thermocellum to
identify metabolic limitations to ethanol production. Frontiers in Energy Research, 8.
204. Holwerda, E. K., Olson, D. G., Ruppertsberger, N. M., Stevenson, D. M., Murphy, S. J. L.,
Maloney, M. I., Lanahan, A. A., et al., (2020). Metabolic and evolutionary responses of
Clostridium thermocellum to genetic interventions aimed at improving ethanol production.
Biotechnology for Biofuels, 13(1), 40. https://doi.org/10.1186/s13068-020-01680-5.
205. Garcia, S., Thompson, A. R., Giannone, R. J., Dash, S., Maranas, C., & Trinh, C. T.,
(2020). Development of a Genome-Scale Metabolic Model of Clostridium thermocellum
and its Applications for Integration of Multi-Omics Datasets and Strain Design. BioRxiv.
206. Singh, N., Mathur, A. S., Gupta, R. P., Barrow, C. J., Tuli, D., & Puri, M., (2018).
Enhanced cellulosic ethanol production via consolidated bioprocessing by Clostridium
thermocellum ATCC 31924. Bioresource Technology, 250, 860–867.
207. Liu, K., Yuan, X., Liang, L., Fang, J., Chen, Y., He, W., & Xue, T., (2019b). Using CRISPR/
Cas9 for multiplex genome engineering to optimize the ethanol metabolic pathway in
Saccharomyces cerevisiae. Biochemical Engineering Journal, 145, 120–126. https://doi.
org/10.1016/j.bej.2019.02.017.
Apple Academic Press
208. Kim, S. K., & Westpheling, J., (2018). Engineering a spermidine biosynthetic pathway
in Clostridium thermocellum results in increased resistance to furans and increased
ethanol production. Metabolic Engineering, 49, 267–274.
209. Cheng, C., Bao, T., & Yang, S. T., (2019). Engineering Clostridium for improved solvent
production: Recent progress and perspective. Applied Microbiology and Biotechnology,
103(14), 5549–5566.
210. Lo, J., Olson, D. G., Murphy, S. J. L., Tian, L., Hon, S., Lanahan, A., & Lynd, L. R.,
(2017). Engineering electron metabolism to increase ethanol production in Clostridium
thermocellum. Metabolic Engineering, 39, 71–79.
Author Copy
211. Rydzak, T., Garcia, D., Stevenson, D. M., Sladek, M., Klingeman, D. M., Holwerda, E.
K., & Guss, A. M., (2017). Deletion of type I glutamine synthetase deregulates nitrogen
metabolism and increases ethanol production in Clostridium thermocellum. Metabolic
Engineering, 41, 182–191.
212. Riederer, A., Takasuka, T. E., Makino, S. I., Stevenson, D. M., Bukhman, Y. V., Elsen, N.
L., & Fox, B. G., (2011). Global gene expression patterns in Clostridium thermocellum
as determined by microarray analysis of chemostat cultures on cellulose or cellobiose.
Applied and Environmental Microbiology, 77(4), 1243–1253.
213. Tian, L., Papanek, B., Olson, D. G., Rydzak, T., Holwerda, E. K., Zheng, T., & Hettich,
R. L., (2016). Simultaneous achievement of high ethanol yield and titer in Clostridium
thermocellum. Biotechnology for Biofuels, 9(1), 1–11.
214. Hon, S., Olson, D. G., Holwerda, E. K., Lanahan, A. A., Murphy, S. J., Maloney,
M. I., & Lynd, L. R., (2017). The ethanol pathway from Thermoanaerobacterium
saccharolyticum improves ethanol production in Clostridium thermocellum. Metabolic
Engineering, 42, 175–184.
215. Deng, Y., Olson, D. G., Zhou, J., Herring, C. D., Shaw, A. J., & Lynd, L. R., (2013).
Redirecting carbon flux through exogenous pyruvate kinase to achieve high ethanol
yields in Clostridium thermocellum. Metabolic Engineering, 15, 151–158.
216. Argyros, D. A., Tripathi, S. A., Barrett, T. F., Rogers, S. R., Feinberg, L. F., Olson, D.
G., & Caiazza, N. C., (2011). High ethanol titers from cellulose by using metabolically
engineered thermophilic, anaerobic microbes. Applied and Environmental Microbiology,
77(23), 8288–8294.
217. Rydzak, T., Lynd, L. R., & Guss, A. M., (2015). Elimination of formate production in
Clostridium thermocellum. Journal of Industrial Microbiology & Biotechnology, 42(9),
1263–1272.
218. Verbeke, T. J., Giannone, R. J., Klingeman, D. M., Engle, N. L., Rydzak, T., Guss, A. M.,
& Elkins, J. G., (2017). Pentose sugars inhibit metabolism and increase expression of an
AgrD-type cyclic pentapeptide in Clostridium thermocellum. Scientific Reports, 7, 43355.
219. Xiong, W., Reyes, L. H., Michener, W. E., Maness, P. C., & Chou, K. J., (2018). Engi-
neering cellulolytic bacterium Clostridium thermocellum to co-ferment cellulose-and
hemicellulose-derived sugars simultaneously. Biotechnology and Bioengineering, 115(7),
1755–1763.
220. Van, D. V. D., Lo, J., Brown, S. D., Johnson, C. M., Tschaplinski, T. J., Martin, M.,
& Guss, A. M., (2013). Characterization of Clostridium thermocellum strains with
Author Copy
gene in Saccharomyces cerevisiae. World Journal of Microbiology and Biotechnology,
34(10), 154. https://doi.org/10.1007/s11274-018-2518-4.
225. Fan, D., Li, C., Fan, C., Hu, J., Li, J., Yao, S., & Luo, K., (2020). MicroRNA6443-
mediated regulation of ferulate 5-hydroxylase gene alters lignin composition and
enhances saccharification in Populus tomentosa. New Phytologist, 226(2), 410–425.
226. Liu, S., Liu, Y. J., Feng, Y., Li, B., & Cui, Q., (2019a). Construction of consolidated bio-
saccharification biocatalyst and process optimization for highly efficient lignocellulose
solubilization. Biotechnology for Biofuels, 12(1), 35.
227. Toivari, M. H., Aristidou, A., Ruohonen, L., & Penttilä, M., (2001). Conversion of
xylose to ethanol by recombinant Saccharomyces cerevisiae: Importance of xylulokinase
(XKS1) and oxygen availability. Metabolic Engineering, 3(3), 236–249.
228. Fu, C., Mielenz, J. R., Xiao, X., Ge, Y., Hamilton, C. Y., Rodriguez, M., ... & Dixon, R.
A. (2011). Genetic manipulation of lignin reduces recalcitrance and improves ethanol
production from switchgrass. Proceedings of the National Academy of Sciences, 108(9),
3803–3808.
229. Bhatia, S. K., Kim, S. H., Yoon, J. J., & Yang, Y. H. (2017). Current status and strategies
for second generation biofuel production using microbial systems. Energy Conversion
and Management, 148, 1142–1156.
230. Javed, U., Ansari, A., Aman, A., & Qader, S. A. U. (2019). Fermentation and saccharifi-
cation of agro-industrial wastes: A cost-effective approach for dual use of plant biomass
wastes for xylose production. Biocatalysis and Agricultural Biotechnology, 21, 101341.
231. Ayodele, B. V., Alsaffar, M. A., & Mustapa, S. I. (2020). An overview of integration
opportunities for sustainable bioethanol production from first-and second-generation
sugar-based feedstocks. Journal of Cleaner Production, 245, 118857.
232. Ledesma-Amaro, R., Lazar, Z., Rakicka, M., Guo, Z., Fouchard, F., Crutz-Le Coq, A.
M., & Nicaud, J. M. (2016b). Metabolic engineering of Yarrowia lipolytica to produce
chemicals and fuels from xylose. Metabolic Engineering, 38, 115–124.
233. Hahn-Hägerdal, B., Galbe, M., Gorwa-Grauslund, M. F., Lidén, G., & Zacchi, G. (2006).
Bio-ethanol–the fuel of tomorrow from the residues of today. Trends in biotechnology,
24(12), 549–556.
Author Copy
SUBBURAMU KARTHIKEYAN2
1
Department of Agricultural Microbiology,
Tamil Nadu Agricultural University, Coimbatore, Tamil Nadu, India
Department of Renewable Energy Engineering, Tamil Nadu Agricultural
2
ABSTRACT
The growing population and increased urbanization have set great demand for
energy. With the deprivation of fossil fuels, interest has focused on renewable
sources. Among them, bioethanol is an attractive and green source being used
as a promising alternative to reduce environmental pollution. In the process of
ethanol production, varied ranges of feedstocks are being converted through
microbial fermentation. Widespread industrial biocatalysts are exploited in
bioethanol production like bacteria, yeast, fungi, and algae according to the type
of raw material, environmental conditions, and resources available. However,
the organisms leading the ethanol industries are Saccharomyces cerevisiae,
Pichia stipites, Zymomonas mobilis, and Clostridium thermocellum. Their
capacity to yield higher ethanol, ability to ferment sugars, growth in simple
and inexpensive media, resistance to inhibitors, and contaminants makes
them potential candidates in the fermentation process. While the maintenance
of growth and metabolic efficiency of these microbes with resistance to
various stresses is the desirable factor, several challenges predominate in the
microbial fermentation inhibiting the overall ethanol production in which
ethanol toxicity is a significant concern. The key advantages of microbial
strain employed for industrial ethanol production are, ethanol tolerance
Author Copy
essential. With the assistance of molecular tools, the strain with improved
ethanol tolerance can be achieved.
6.1 INTRODUCTION
Author Copy
pH preference 2.0–6.5 3.5–7.5 6.0–8.5
Metabolic pathway EMP ED ABE fermentation
Theoretical ethanol yield >90% >90% 2.5%
Ethanol tolerance (w/v) 15% 16% 1–2%
Safety GRAS GRAS GRAS
Author Copy
tolerant strain depicting its potential mechanism for stress resistance [7].
tolerance. This necessitates the need for ethanol tolerant strains which could
be achieved through several reprogramming strategies.
Apple Academic Press
TABLE 6.2 Effect of Ethanol on the Physiology and Metabolism of Fermenting Microorganisms
Effect References
S. cerevisiae At low concentration: [17]
Inhibits cell division, growth rate and decreases cell volume
At high concentration:
Decreases cell vitality and causes death
Function of ribosome is inhibited [18]
Author Copy
Governs vitality and viability of the cells, thereby terminating [19]
fermentation; Inhibits glucose and amino acids transport
systems, denaturation key glycolytic enzymes
Membrane permeabilization, cytosolic, and vacuolar [20]
acidification
Dissipation of membrane integrity and increase in [21]
membrane permeability due to intercalation of lipid bilayer
Z. mobilis Decreased cell mass and ethanol production, Inhibition of [22]
fermentation process
Decreased rate of substrate conversion [23]
Inactivation of enzymes involved in ethanol production [24]
C. thermocellum Imbalance in NADPH [3]
Interruption in proton motive force and ATP production [25]
Few studies revealed the genes involved in ethanol stress guiding the
rational strategies to progress the process performance. The genome sequence
of Z. mobilisZM4 revealed the sigma factor (σE, ZMO4104) plays a crucial
role in against ethanol stress [12]. Alcohol dehydrogenase II (ADH2) protein,
encoded by the adhB gene was found to be the major protein involved in
ethanol stress [13]. Further Pallach et al. [14] have proposed the exopoly-
saccharides (EPS) components of Z. mobilis contributes to its high ethanol
tolerance mechanism.
Ethanol fermentation using thermophilic bacteria has been suggested as
one of the novel systems. However, the membrane composition of S. cerevi-
siae and Z. mobilis provides higher tolerance to ethanol. C. thermocellum is
a gram-positive, thermophilic anaerobe that serves as an efficient candidate
in CBP due to its ability to rapidly fermenting cellulosic biomass. One of
the challenges in the industrial application hindering their potential is low
ethanol yield, titer, and tolerance producing acetate, lactate, H2, formate as
Author Copy
partitioning into a lipid bilayer. The aliphatic chains of ethanol get deeply
inserted into the hydrophobic interior of a lipid bilayer and favor a high
degree of permeability. As a tolerance mechanism, the cell wall lipid
composition is increased with polyunsaturated fatty acids, ergosterol, and
phosphatidylcholine. The cell wall architecture gets remodeled to attain
robust nature through various signaling pathways [26]. The increased
membrane permeability also results in cytosolic and vacuolar acidification
for which vacuolar H+-ATPase (V-ATPase) plays a crucial role in the
maintenance of intracellular pH homeostasis. The intracellular transport of
H+ into vacuoles by H+ V-ATPase assists in translocation across the vacuolar
membrane through hydrolysis of ATP, thereby counteracts the ethanol stress.
Further several genes would up regulate upon ethanol stress leading to an
alteration in the transcription machinery which renders protection to the
organism by various physiological changes. Therefore, the design of the
ethanol tolerant organisms can be concentrated on the above-mentioned
background with numerous techniques as follows. The creation of strains
with augmented stress tolerance is however attainable through a range of
approaches (Figure 6.2), combining with recent technology could limit time
and tedious multistep. Traditional approaches for the development of ethanol
tolerant phenotype include rational selection approaches for modification
followed by probability-driven processes like adaptive evolution or random
mutagenesis. The strain development for ethanol tolerance has certain
gold standards like mutagenesis and selection traditionally in the industry.
The advent of recombinant tools gives rise to more sophisticated alternate
strategies to ameliorate alcohol toxicity. The advanced omics technologies
such as genomics, proteomics, transcriptomics, and metabolomics have
aided to understand the complex alcohol toxicity and tolerating mechanism.
Exploring the molecular basis of ethanol tolerance en routes the development
of rational approaches. The metabolomic analyzes of E. coli on the responses
to ethanol and butanol tolerance elucidated several amino acids like valine,
glycine, isoleucine, glutamic acid and osmoprotectants like trehalose. Most
transcriptomic analyzes have shown the role of small molecular weight
compounds on alcohol tolerance.
Apple Academic Press
FIGURE 6.1 Ethanol tolerance mechanism of mentioned microbes. (a) change in membrane Author Copy
composition; (b) maintenance of intracellular pH by H+ ATPase pump; (c and d) production of
stress-responsive proteins/genes; (e) vacuole acidification/vacuole mediated transport.
biosynthesis genes FKS2, CRH1, and PIR3 thereby causes remodeling of the
cell wall [28]. The mutants lacking CWI genes were found to be sensitive to
ethanol.
Apple Academic Press
Author Copy
FIGURE 6.2 Various approaches are available to develop ethanol tolerant strains.
Author Copy
(sir) proteins, and aromatic amino acids (AAA) helping in the survival of cells.
Author Copy
with improved biomass and cell viability
Supplementation of Mg2+ in the 50% cells remained viable in exposure of [38]
culture media 20% (v/v) ethanol for 9 h
Exponential feeding of vitamins Increase in final ethanol titer and ethanol [39]
tolerance
Supplementation of CaCl2 and Protection against toxic effect of ethanol [40]
MgCl2
Non-limiting oxygen condition 23% increase in viable cell mass [41]
Supplementation of 10–20 mM Reduction in cell mortality [22]
magnesium
Author Copy
strain. Selective pressure of inhibitory concentration of hexanol evolved
tolerant S. cerevisiae BY4741 [49]. The sequential transfer method with
increasing ethanol concentration in Z. mobilis ZM4 mutant ER7ap exhibited
better performance than wild type [50].
Author Copy
tion is screened to identify the best performing isolate. Third, the genomes
of supreme cells are shuffled either by sexual or asexual means following
protoplast fusion or sporulation, respectively, in single strain recursive
protoplast fusion between multi-parent strains.
In the past years, novel genome editing tools like transcription activator-like
effector nuclease (TALEN's), zinc finger nuclease (ZFN) have been used to
engineer strains with desirable traits in industrial settings. However, there are
only a few studies applying the gene-editing tools for ethanol tolerance. Zhang
et al. [58] have demonstrated TALEN assisted multiplex editing (TAME) for
improving ethanol tolerance in yeast by multiple genetic perturbations in the
genome. Mitsui, Yamada, and Ogino [59] have achieved higher cell viability
in low pH and high ethanol concentration in S. cerevisiae using Clustered
regularly interspaced short palindromic repeats (CRISPR)-CRISPR associ-
ated protein (Cas). The large-scale rearrangement of genomic DNA can be a
Apple Academic Press
6.3.4.4 OVER-EXPRESSION
The iron-sulfur clusters (ISC) are the prosthetic groups responsible for
catalytic and regulatory functions found in both prokaryotes and eukaryotes.
Author Copy
The components of the mitochondrial ISC in S. cerevisiae were reported to
be related in iron homeostasis due to ethanol tolerance and are encoded by
the genes NFS1, ISU1, ISU2, ISA1, ISA2, JAC1, SSQ1, YAH1, GRX5, and
IBA5. The over-expression of Jac 1p and Isu 1p in S. cerevisiae UMArn3 was
evaluated and revealed their counteracting ability of metabolic unbalance
caused by an increase in ROS generation [60]. Further Snf complex of S.
cerevisiae, a member of AMP protein kinase family engages in a series of
cellular functions and responds to environmental stresses. The effect of Snf
over-expression revealed a 39% increase in cell survival rate over parental
strain at 8% ethanol accompanied with altered expression of genes involved
in glucose transport and accumulation of fatty acids [31].
epPCR improved glucose consumption and ethanol yield under ethanol stress
in Z. mobilis [54]. Further, incorporation of gene modules increased fatty acid
production which served a crucial criterion for ethanol tolerance [62].
Beyond yeasts being physiologically ideal, the complex array of stress
factors exhibited have to be encountered cooperatively. Conventional breeding
techniques bestow bounded up-gradation of strain robustness. However,
their painstaking standards assist “omics” technologies in recent research.
Author Copy
Exploring genetic elements required to maintain ethanol tolerance would be
a better approach rather than genetic modifications. On the other hand, altera-
tion in the gene for ethanol tolerance could compromise several fermentation
parameters. The ever-expanding knowledge on yeast genome and functions
makes an obscure platform for commercial-scale ethanol tolerant strains.
Recent precise technology for genome editing enabled quick engineering of
yeast strains which would develop strains for numerous applications.
Author Copy
NADPH. In C. thermocellum, the role of adhE is crucial and their deletion
strain showed >90% activity loss of ALDH, ADH, and >95% reduction of
ethanol production [67]. Point mutation resulted in the replacing certain
amino acids at particular sites of ADH [3].
The genome of six isolates with n-butanol tolerance was re-sequenced and
the mutations in the coding sequence of Clo1313_0853 and Clo1313_1798
were studied [5]. The gene Clo1313_0853 catalyzes the hydrolysis of
phosphatidylcholine and phospholipids to produce phosphatidic acid (PA).
The mutation in the gene truncates the protein by frameshift and protects
the membrane in the presence of n-butanol/ethanol. Mutation in the gene
Clo1313_1798, encoding alcohol dehydrogenase (adhE) increases its
ability to use NADPH as a co-factor yielding increased ethanol and titer.
The increased NADPH activities are associated with AdhE mutant, where
the reverse flux through ADH reaction affects NADH/NAD+ ratio and the
NADPH/NADP+ ratio. Deletion of adhE strain (LL1111) increases tolerance
and inhibits several primary alcohols by reverse flux through AdhE [68].
6.4 CONCLUSION
resourcefully with the aid of synthetic biology and omics techniques. The
ethanol-related industries may take advantage of these findings to lower the
cost of ethanol production.
Apple Academic Press
KEYWORDS
• bioethanol
• Clostridium thermocellum
• ethanol tolerance
Author Copy
• exopolysaccharides
• yeast
• Zymomonas mobilis
REFERENCES
1. Yang, S., Fei, Q., Zhang, Y., Contreras, L. M., Utturkar, S. M., Brown, S. D., Himmel,
M. E., & Zhang, M., (2016). Zymomonas mobilis as a model system for production of
biofuels and biochemicals. Microb. Biotechnol., 9, 699–717.
2. Gunasekaran, P., Karunakaran, T., & Kasthuribai, M., (1986). Fermentation pattern of
Zymomonas mobilis strains on different substrates—a comparative study. J. Biosci., 10,
181–186.
3. Brown, S. W., & Oliver, S. G., (1982). Isolation of ethanol-tolerant mutants of yeast by
continuous selection. Eur. J. Appl. Microbiol. Biotechnol., 16, 119–122.
4. Rani, K. S., & Seenayya, G., (1999). High ethanol tolerance of new isolates of
Clostridium thermocellum strains SS21 and SS22. World J. Microbiol. Biotechnol., 15,
173–178.
5. Shao, X., Raman, B., Zhu, M., Mielenz, J. R., Brown, S. D., Guss, A. M., & Lynd, L. R.,
(2011). Mutant selection and phenotypic and genetic characterization of ethanol-tolerant
strains of Clostridium thermocellum.Appl. Microbiol. Biotechnol., 92, 641–652.
6. Timmons, M. D., Knutson, B. L., Nokes, S. E., Strobel, H. J., & Lynn, B. C., (2009).
Analysis of composition and structure of Clostridium thermocellum membranes from
wild-type and ethanol-adapted strains. Appl. Microbiol. Biotechnol., 82, 929–939.
7. Zhu, X., Cui, J., Feng, Y., Fa, Y., Zhang, J., & Cui, Q., (2013). Metabolic adaption of
ethanol-tolerant Clostridium thermocellum. PLoS One, 8, e70631.
8. Albano, E., (2007). Alcohol, oxidative stress and free radical damage. Proc. Nutr. Soc.,
65, 278–290.
9. Ding, J., Huang, X., Zhang, L., Zhao, N., Yang, D., & Zhang, K., (2009). Tolerance
and stress response to ethanol in the yeast Saccharomyces cerevisiae. Appl. Microbiol.
Biotechnol., 85, 253–263.
10. Rupcic, J., & Juresic, G. C., (2010). Influence of stressful fermentation conditions
on neutral lipids of a Saccharomyces cerevisiae brewing strain. World J. Microbiol.
Biotechnol., 26, 1331–1336.
11. Hahn-Hägerdal, B., Galbe, M., Gorwa-Grauslund, M. F., Lidén, G., & Zacchi, G., (2006).
Apple Academic Press
Bio-ethanol - the fuel of tomorrow from the residues of today. Trends Biotechnol., 24,
549–556.
12. Seo, J. S., Chong, H., Park, H. S., Yoon, K. O., Jung, C., Kim, J. J., Hong, J. H., et
al., (2005). The genome sequence of the ethanologenic bacterium Zymomonas mobilis
ZM4. Nat. Biotechnol., 23, 63–68.
13. An, H., Scopes, R. K., Rodriguez, M., Keshav, K. F., & Ingram, L. O., (1991). Gel
electrophoretic analysis of Zymomonas mobilis glycolytic and fermentative enzymes:
Identification of alcohol dehydrogenase II as a stress protein. J. Bacteriol., 173, 5975–5982.
14. Pallach, M., Marchetti, R., Di Lorenzo, F., Fabozzi, A., Giraud, E., Gully, D., Paduano,
Author Copy
L., et al., (2018). Zymomonas mobilis exopolysaccharide structure and role in high
ethanol tolerance. Carbohydr. Polym., 201, 293–299.
15. Ellis, L. D., Holwerda, E. K., Hogsett, D., Rogers, S., Shao, X., Tschaplinski, T., Thorne,
P., & Lynd, L. R., (2012). Closing the carbon balance for fermentation by Clostridium
thermocellum (ATCC 27405). Bioresour Technol., 103, 293–299.
16. Biswas, R., Zheng, T., Olson, D. G., Lynd, L. R., & Guss, A. M., (2015). Elimination of
hydrogenase active site assembly blocks H2 production and increases ethanol yield in
Clostridium thermocellum. Biotechnol. Biofuels, 8, 1–8.
17. Birch, R. M., & Walker, G. M., (2000). Influence of magnesium ions on heat shock and
ethanol stress responses of Saccharomyces cerevisiae. Enzyme Microb. Technol., 26,
678–687.
18. Li, R., Miao, Y., Yuan, S., Li, Y., Wu, Z., & Weng, P., (2019). Integrated transcriptomic
and proteomic analysis of the ethanol stress response in Saccharomyces cerevisiae
Sc131. J. Proteomics, 203, 103377.
19. Stanley, D., Bandara, A., Fraser, S., Chambers, P. J., & Stanley, G. A., (2010). The ethanol
stress response and ethanol tolerance of Saccharomyces cerevisiae. J. Appl. Microbiol.,
109,13–24.
20. Charoenbhakdi, S., Dokpikul, T., Burphan, T., Techo, T., & Auesukaree, C., (2016).
Vacuolar H+ ATPase protects Saccharomyces cerevisiae cells against ethanol-induced
oxidative and cell wall stresses. Appl. Environ. Microbiol., 82, 3121–3130.
21. Weber, F. J., & De Bont, J. A. M., (1996). Adaptation mechanisms of microorganisms to
the toxic effects of organic solvents on membranes. BBA. Rev. Biomembr., 1286, 225–245.
22. Thanonkeo, P., Laopaiboon, P., Sootsuwan, K., & Yamada, M., (2007). Magnesium ions
improve growth and ethanol production of Zymomonas mobilis under heat or ethanol
stress. Biotechnol., 6, 112–119.
23. Laudrin, I., & Goma, G., (1982). Ethanol production by Zymomonas mobilis: Effect of
temperature on cell growth, ethanol production and intracellular ethanol accumulation.
Biotechnol. Lett., 4, 537–542.
24. Millar, D. G., Griffiths-Smith, K., Algar, E., & Scopes, R. K., (1982). Activity and
stability of glycolytic enzymes in the presence of ethanol. Biotech. Lett., 4, 601–606.
25. Demain, A. L., Newcomb, M., & Wu, J. H. D., (2005). Cellulase, clostridia, and ethanol.
Microbiol. Mol. Biol. Rev., 69,124–154.
26. Levin, D. E., (2011). Regulation of cell wall biogenesis in Saccharomyces cerevisiae:
The cell wall integrity signaling pathway. Genetics, 189(4), 1145–1175.
27. Ma, M., & Liu, Z. L., (2010). Mechanisms of ethanol tolerance in Saccharomyces
cerevisiae. Appl. Microbiol. Biotechnol., 87, 829–845.
28. Udom, N., Chansongkrow, P., Charoensawan, V., & Auesukaree, C., (2019). Coordination
of the cell wall integrity and high-osmolarity glycerol pathways in response to ethanol
Apple Academic Press
Author Copy
of the SNF1 gene and varied beta isoform of Snf1 dominates in stresses. Microbial Cell
Factories, 19, 134.
32. Ansarypour, Z., & Shahpiri, A., (2017). Heterologous expression of a rice metallothionein
isoform (OsMTI-1b) in Saccharomyces cerevisiae enhances cadmium, hydrogen
peroxide and ethanol tolerance. Braz. J. Microbiol., 48, 537–543.
33. Zyrina, A. N., Smirnova, E. A., Markova, O. V., Severin, F. F., & Knorre, D. A., (2017).
Mitochondrial superoxide dismutase and Yap1p Act as a signaling module contributing
to ethanol tolerance of the yeast Saccharomyces cerevisiae. Appl. Environ. Microbiol.,
83, e02759–02716.
34. Li, R., Xiong, G., Yuan, S., Wu, Z., Miao, Y., & Weng, P., (2017). Investigating the
underlying mechanism of Saccharomyces cerevisiae in response to ethanol stress
employing RNA-seq analysis. World J. Microbiol. Biotechnol., 33, 206.
35. Xue, C., Zhao, X. Q., Yuan, W. J., & Bai, F. W., (2008). Improving ethanol tolerance of
a self-flocculating yeast by optimization of medium composition. World J. Microbiol.
Biotechnol., 24, 2257.
36. Ahmed, K., Zhang, M., Wang, M., & Wang, C., (2020). Effects of zinc combining with
specific metal ions on ethanol tolerance of yeast Saccharomyces cerevisiae. Int. J. Agric.
Biol., 23, 566–572.
37. Xu, Y., Yang, H., Brennan, C. S., Coldea, T. E., & Zhao, H., (2020). Cellular mechanism
for the improvement of multiple stress tolerance in brewer’s yeast by potassium ion
supplementation. Int. J. Food Sci Technol., 55, 2419–2427.
38. Hu, C. K., Bai, F. W., & An, L. J., (2003). Enhancing ethanol tolerance of a self-
flocculating fusant of Schizosaccharomyces pombe and Saccharomyces cerevisiae by
Mg2+ via reduction in plasma membrane permeability. Biotechnol. Lett., 25:1191–1194.
39. Alfenore, S., Molina, J. C., Guillouet, S., Uribelarrea, J. L., Goma, G., & Benbadis,
L., (2002). Improving ethanol production and viability of Saccharomyces cerevisiae by
a vitamin feeding strategy during fed-batch process. Appl. Microbiol. Biotechnol., 60,
67–72.
40. Ciesarova, Z., Smogrovicova, D., & Domeny, Z., (1996). Enhancement of yeast ethanol
tolerance by calcium and magnesium. Folia Microbiol., 41, 485–488.
41. Alfenore, S., Cameleyre, X., Benbadis, L., Bideaux, C., Uribelarrea, J. L., Goma,
G., Molina, J. C., & Guillouet, S. E., (2004). Aeration strategy: A need for very high
ethanol performance in Saccharomyces cerevisiae fed-batch process. Appl. Microbiol.
Biotechnol., 63, 537–542.
42. Dinh, T. N., Nagahisa, K., Hirasawa, T., Furusawa, C., & Shimizu, H., (2008). Adaptation
of Saccharomyces cerevisiae cells to high ethanol concentration and changes in fatty
acid composition of membrane and cell size. PLoS One, 3, e2623.
43. Fiedurek, J., Skowronek, M., & Gromada, A., (2011). Selection and adaptation of
Apple Academic Press
Author Copy
tolerance of an ethanologenic Saccharomyces cerevisiae strain by freeze-thaw treatment.
Biotechnol. Lett., 29, 1501–1508.
47. Morard, M., Macías, L. G., Adam, A. C., Lairón-Peris, M., Pérez-Torrado, R., Toft, C.,
& Barrio, E., (2019). Aneuploidy and ethanol tolerance in Saccharomyces cerevisiae.
Front. Genet., 10, 82.
48. Turanli-Yildiz, B., Benbadis, L., Alkım, C., Sezgin, T., Akşit, A., Gökçe, A., Oztürk, Y.,
et al., (2017). In vivo evolutionary engineering for ethanol-tolerance of Saccharomyces
cerevisiae haploid cells triggers diploidization. J. Biosci. Bioeng., 124, 309–318.
49. Davis, L. S. A., Griffith, D. A., Choi, B., Cate, J. H. D., & Tullman-Ercek, D.,
(2018). Evolutionary engineering improves tolerance for medium-chain alcohols in
Saccharomyces cerevisiae. Biotechnol. Biofuels, 11(1), 90.
50. Carreón-Rodríguez, O. E., Gutiérrez-Ríos, R. M., Acosta, J. L., Martinez, A., &
Cevallos, M. A., (2019). Phenotypic and genomic analysis of Zymomonas mobilis ZM4
mutants with enhanced ethanol tolerance. Biotechnol. Rep., 23, e00328.
51. Ramírez-Cota, G. Y., López-Villegas, E. O., Jiménez-Aparicio, A. R., & Hernández-
Sánchez, H., (2020). Modeling the ethanol tolerance of the probiotic yeast Saccharomyces
cerevisiae var. boulardii CNCM I-745 for its possible use in a functional beer. Probiotics
Antimicrob. Proteins., 1–8.
52. Brenac, L., Baidoo, E. E. K., Keasling, J. D., & Budin, I., (2019). Distinct functional
roles for hopanoid composition in the chemical tolerance of Zymomonas mobili. Mol.
Microbiol., 112, 1564–1575.
53. Shi, D. J., Wang, C. L., & Wang, K. M., (2009). Genome shuffling to improve thermo-
tolerance, ethanol tolerance and ethanol productivity of Saccharomyces cerevisiae. J. Ind.
Microbiol. Biotechnol., 36, 139–147.
54. Jetti, K. D., Gns, R. R., Garlapati, D., & Nammi, S. K., (2019). Improved ethanol
productivity and ethanol tolerance through genome shuffling of Saccharomyces
cerevisiae and Pichia stipitis. Int Microbiol., 22, 247–254.
55. Hou, L., (2010). Improved production of ethanol by novel genome shuffling in
Saccharomyces cerevisiae. Appl. Biochem. Biotechnol., 160, 1084–1093.
56. Snoek, T., Picca, N. M., Van, D. B. S., Mertens, S., Saels, V., Verplaetse, A., Steensels,
J., & Verstrepen, K. J., (2015). Large-scale robot-assisted genome shuffling yields
industrial Saccharomyces cerevisiae yeasts with increased ethanol tolerance. Biotechnol.
Biofuels, 8, 32.
57. Xin, Y., Yang, M., Yin, H., & Yang, J., (2020). Improvement of Ethanol Tolerance by
Inactive Protoplast Fusion in Saccharomyces cerevisiae. Biomed Res Int., 1979318.
58. Zhang, G., Lin, Y., Qi, X., Li, L., Wang, Q., & Ma, Y., (2015). TALENs-assisted multiplex
editing for accelerated genome evolution to improve yeast phenotypes. ACS Synth. Biol.,
Apple Academic Press
4, 1101–1111.
59. Mitsui, R., Yamada, R., & Ogino, H., (2019). Improved stress tolerance of Saccharomyces
cerevisiae by CRISPR-Cas-mediated genome evolution. Appl. Biochem. Biotechnol.,
189, 810–821.
60. Martínez-Alcántar, L., Madrigal, A., Sánchez-Briones, L., Díaz-Pérez, A. L., López-
Bucio, J. S., & Campos-García, J., (2019). Over-expression of Isu1p and Jac1p increases
the ethanol tolerance and yield by superoxide and iron homeostasis mechanism in an
engineered Saccharomyces cerevisiae yeast. J. Ind. Microbiol. Biotechnol., 46, 925–936.
61. Warfield, L., Ramachandran, S., Baptista, T., Devys, D., Tora, L., & Hahn, S., (2017).
Author Copy
Transcription of nearly all yeast RNA polymerase II-transcribed genes is dependent on
transcription factor TFIID. Mol. Cell, 68, 118–129.e115.
62. Wang, H., Cao, S., Wang, W. T., Wang, K. T., & Jia, X., (2016). Very high gravity ethanol
and fatty acid production of Zymomonas mobilis without amino acid and vitamin. J. Ind.
Microbiol. Biotechnol., 43, 861–871.
63. Cho, S. H., Haning, K., & Contreras, L. M., (2015). Strain engineering via regulatory
noncoding RNAs: Not a one-blueprint-fits-all. Curr. Opin. Chemi. Eng., 10, 25–34.
64. Cho, S. H., Lei, R., Henninger, T. D., & Contreras, L. M., (2014). Discovery of ethanol-
responsive small RNAs in Zymomonas mobilis. Appl. Environ. Microbiol., 80,4189–4198.
65. Han, R., Haning, K., Gonzalez-Rivera, J. C., Yang, Y., Li, R., Cho, S. H., Huang, J., et
al., (2020). Multiple small RNAs interact to co-regulate ethanol tolerance in Zymomonas
mobilis. Front. Bioeng. Biotechnol., 8, 155.
66. Haning, K., Engels, S. M., Williams, P., Arnold, M., & Contreras, L. M., (2020).
Applying a new REFINE approach in Zymomonas mobilis identifies novel sRNAs that
confer improved stress tolerance phenotypes. Front. Microbiol., 10, 2987.
67. Lo, J., Zheng, T., Hon, S., Olson, D. G., & Lynd, L. R., (2015). The bifunctional
alcohol and aldehyde dehydrogenase gene, adhE, is necessary for ethanol production in
Clostridium thermocellum and Thermoanaerobacterium saccharolyticum. J. Bacteriol.,
197, 1386–1393.
68. Tian, L., Cervenka, N. D., Low, A. M., Olson, D. G., & Lynd, L. R., (2019). A mutation
in the AdhE alcohol dehydrogenase of Clostridium thermocellum increases tolerance to
several primary alcohols, including isobutanol, n-butanol and ethanol. Sci. Rep., 9, 1736.
Author Copy
Department of Bioscience and Biotechnology, Banasthali Vidyapith,
Rajasthan–304022, India, Tel.: 9910331924,
E-mail: charusaraf17@gmail.com (C. Saraf), Tel.: 9887398384,
E-mail: kakoli_dutt@rediffmail.com (K. Dutt)
ABSTRACT
The ability to harness fire from burning wood was the first application of fuel-
based energy. For a long time, wood and other derived products were the sources
of energy to support various activities like cooking, metallurgy, pottery, etc.
The discovery of fossil fuels and the development of the steam engine opened
a new age of applications. But in this modern world, overuse of fossil fuels to
sustain various technologies has started to take a heavy toll in the form of envi-
ronmental pollution. Thus, the new search for alternative energy sources which
can efficiently replace the fossil fuels and have minimal environmental issues
has become mandatory, with several biofuels such as bioethanol and biodiesel
gaining prominence. Ethanol production, a massive global commercial venture
shows continuous evolution in the production process for yield enhancement.
The developments observed in the characterization of different biomass as raw
material for ethanol production has led to their categorization as first, second,
third, and fourth-generation fuels. Numerous reviews and research articles
account for the advantages and disadvantages of these substrates against each
other. Additionally, the pretreatment and the fermentation processes play a pivotal
role in the final yield obtained. Thus, in this chapter, an attempt has been made
to present a comprehensive summary of ethanol production with emphasis on
the different substrates, pretreatment, and fermentation processes. In the chapter,
bottleneck of the bioethanol commercialization has also been briefly discussed.
7.1 INTRODUCTION
Over six decades ago, petroleum products were inarguably the source of energy
that kept the world functional and expanding. Nevertheless, in the beginning
Apple Academic Press
Author Copy
quickly became a part of fuel industry as blended fuel ‘gasohol’ [5].
“Fuel for future” a term coined by Henry Ford for ethanol is a major
contender among several key alternative energy sources. Obtained either
from biomass through microbial fermentation or as a byproduct of several
industrial processes [6], ethanol shows valid promise not only as feedstock
for chemical industries but also as biofuel (Figure 7.1). The oxygenated
nature of alcohol enhances its combustion efficiency which added to the
properties of eco-friendly, less toxicity, higher heat of vaporization and
flexibility of sugar substrate for alcohol fermentation increases its appeal
[7–10]. Ethanol has higher octane number (106–110) than gasoline (91–96),
resulting in enhancement of gasoline performance on blending [11–13].
Although, literature reported shows unlimited benefits of bioethanol with
respect to environment and economy, its global development as a fuel and
chemical feedstock is facing a serious bottle neck of usable and fermentable
substrates [14, 15]. Significant researches are going on which are globally
supported by private conglomerates, think tanks, NGOs, and Government
funded projects. This chapter provides an overview of the evolution of
ethanol as a fuel, fermentable substrates, technological approaches, and
fermentation conditions for bioethanol formation with a brief outline on the
hurdles in bioethanol commercialization.
Uses of ethanol as fuel was first reported in 1826 in America and subse-
quently after 50 decades by Nikolaus Otto in 1876, who used ethanol to
power an early internal combustion engine (Figure 7.2). Before the American
Civil War, ethanol was used for lighting but this was severely curtailed due
to liquor tax imposition. After the tax was repealed, ethanol resumed its
function as fuel. In fact, Henry Ford's Model T in 1908 was designed to run
on either gasoline or pure alcohol (adapted from: www.ott.doe.gov/biofuels/
history.html).
Apple Academic Press
Author Copy
FIGURE 7.1 Various applications of ethanol.
Prior to the Second World War, industrial use ethanol was produced by
fermenting blackstrap molasses, a cheap, easy to handle, most economical
and easily available substrate. After the war, the low-priced easy avail-
ability of petroleum-based fuel and natural gas curbed the interest in use of
Apple Academic Press
agricultural crops for fuel production which wiped out the economic incen-
tives for production of liquid fuels from crops leading to the generation of
a significant lacuna. Since, Government interest was lost, many wartime
distillers were dismantled and others were converted to beverage alcohol
plants [16].
The revival of alcohol-based fuel started in 1970s, when the supply of
oil started being disrupted due to instability in the Middle East. The major
Author Copy
oil companies of America began to market ethanol as a gasoline extender
and octane booster (adapted from: www.ott.doe.gov/biofuels/history.html).
Gasohol is a mixture of 9 volumes of gasoline and only 1 volume of ethanol.
The total ethanol production in the USA has increased from 0.175 billion
gallons in 1980 to 15.8 billion gallons in 2017 (adapted from: www.afdc.
energy.gov/data/). In 2015–2016, Brazil had produced 8 billion gallons of
ethanol and sold gasoline with a blend of 18 to 27.5% ethanol [17]. Among
the global players, the United States, Brazil, and European Union (EU) are
the major producers of ethanol followed by China among the developing
countries, see Figure 7.3.
Sugar rich substrates are required to produce alcohol and can be categorized
into four generations [18–22]. Since alcohol is produced via the breakdown
Apple Academic Press
Author Copy
1GLUCOSE + 2ATP 2 ETHANOL + 2CO2 + 2ADP (1)
3XYLOSE + 5ADP + 5Pi 5ETHANOL + 5CO2 + 5ATP + 5H2O (2)
Cash crops like cereals and sugar crops were identified as readily ferment-
able substrates for first-generation ethanol (Table 7.1). It has carbon dioxide
(CO2) benefits and is commercially available even today. Sugar crops like
sugarcane and sugar beet contribute 40% of the ethanol production in the
tropical areas like India, Brazil, and Colombia, and the remaining 60% is
contributed by the starch crops which is used by the United States, EU, and
China [27, 50–52]. High corn and sugarcane harvests lead to an increase in
global biofuel production by 9% in 2014, accounting for 74% (of the total)
of ethanol followed by 23% biodiesel (adapted from: http://www.ren21.net/
wpcontent/uploads/2015/07/REN12-GSR2015_Onli nebook_low1.pdf).
Before 1750, sugar production was limited to sugar cane (Saccharum offici-
narum) grown in the tropical and subtropical regions and shipped globally
[53]. During 1880, a German chemist, Andreas Marggraf extracted sugar
from beets (Beta vulgaris) and sugar beet soon replaced sugar cane as the
main source of sugar in continental Europe as it was found that sugar beet
contains enough amounts (16–20%) of sucrose than sugar cane [54, 55]. The
other sources of sugar production are included in Table 7.2.
Author Copy
Challenges in Developing Sustainable Fermentable Substrate 167
Author Copy
mediates and byproducts like molasses. Another byproduct of the industry,
i.e., sugar beet pulp (SBP), is a potential feedstock for biofuels. It contains
20–25% cellulose, 25–36% hemicellulose, 20–25% pectin, 10–15% protein,
and 1–2% lignin content on a dry weight basis [55, 57].
Sugar derived from plant biomass is a mixture of both hexoses and pentoses
represented mainly by glucose and xylose, respectively. Since the wild-type
strains of Saccharomyces cerevisiae do not metabolize xylose, researchers
have developed two approaches to increase the fermentation yields of ethanol
derived from sugar biomass. The first approach via genetic engineering is to
add pentose metabolic pathways in yeast and other natural ethanologens. The
second approach is to improve the ethanol production by modifying microor-
ganisms having the ability to ferment both hexoses and pentoses [58, 59]. A
general procedure for ethanol production from sugar biomass is illustrated in
Figure 7.4. After fermentation, distillation is done which generate is vinasse
as a byproduct, which can be used as an animal feed or as a fertilizer [56].
Grain crops (corn, barley, wheat, or grain sorghum) and tuber crops (cassava,
potato, sweet potato, Jerusalem artichoke, cactus, or arrowroot) contain large
quantities of starch [60]. Isolated native starch from different sources can be
used for further conversion into bio-based products or bioethanol production.
The residue from starch isolation contains proteins and fiber, which has great
potential for application in food and feed production (adapted from: http://
www.bmbf.de/pub/Roadmap_Biorefineries_eng.pdf.). USA is the biggest
corn starch producing countries with 80% of the worldwide market. There,
95% of ethanol is produced from corn, and the rest from barley, wheat, whey,
and beverage residues [61]. Cassava in Thailand was also under investiga-
tion for bioethanol production. Cassava tubers contain by mass 80% starch
and less than 1.5% proteins. Pretreatment of cassava tubers include cleaning,
peeling, chipping, and drying. The dried cassava chips are used for bioethanol
production [62]. There are two established procedures for processing starch
biomass summarized in Table 7.3; Figures 7.5 and 7.6.
Apple Academic Press
Author Copy
FIGURE 7.4 Diagram of ethanol production from sugar biomass [56].
Author Copy
FIGURE 7.5 Diagram for dry-grind ethanol production from starch crops [65].
FIGURE 7.6 Diagram of wet-mill ethanol production from starch biomass [64].
The first-generation biofuels possessed two major issues, firstly, fuel vs. food
trade-off, and secondly, biofuels potential of CO2 reduction (biofuels produc-
tion could release more carbon than CO2 sequestration in the feedback growth
process) [66]. This trade-off situation would affect the producers, distributors,
Apple Academic Press
related markets, and finally, the regional and national economies [67–69].
The use of food crops for ethanol production became debatable due to multiple
reasons; the lignocellulosic biomass (LCB) (Figure 7.7 and Table 7.1) came
Author Copy
into prominence as feedstock for the second-generation bioethanol [71–73].
Biomass of various agricultural wastes like rice husks, corn stalks, wheat, and
barley straws, etc., and forest biomass such as grasses, wood, etc., are placed in
this category [74, 75]. Some industrial wastes such as brewer's spent grains and
spent grains from distillers, municipal solid wastes such as food waste, kraft
paper and paper sludge containing cellulose were also considered [76–79].
They are inexpensive, abundant and non-competitive with the food crops
[80]. However, an additional pretreatment step apart from the conventional
three-step strategy (Figure 7.8), is required to break down the lignin present in
the lignocellulosic material for subsequent hydrolysis and fermentation [81].
By using a suitable pretreatment method [82], the biomass is degraded into
cellulose, hemicelluloses, and lignin polymer which can be further degraded
into byproducts like furfurals, furans, acetic acid, etc. (Figure 7.9).
In general, recognition of new trending knowledge of energy-based
biofuels (compared to fossil fuels) spurred the doubt of economic efficiency
of biofuels [84]. As instance, cellulosic biofuel production is highly energy-
intensive, which means the energy contained in this type of biofuels is lower
than the energy required for its production [85]. Researchers in the past
decades have attempted many experiments with limited success to lower
the cellulosic ethanol production costs [66]. Study using microbial or fungal
systems for more effective and faster cellulose breakdown and fermentation
process has also been attempted [86, 87]. However, no wide-scale commer-
cial solution has been found and the development in this field is still ongoing.
for production. The three generations of biofuels have been compared with
the petroleum products in Table 7.4. In a report by Kim and Day [89], the
Louisiana sugar mills in the USA operate only quarterly as sugar cane avail-
ability is only from October to December, making capital investments for
Apple Academic Press
Author Copy
Author Copy
TABLE 7.4
Apple Academic Press
(Continued)
Author Copy
174 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
Microalgae are the oldest living organisms on Earth [96] which can grow
100-fold faster than terrestrial and can double their biomass in less than one
day time period [97]. It is beneficial as microalgae can grow faster and fixes
the CO2 at a higher rate than terrestrial plants. Lignin is absent in microalgae
making it easier to convert their starch and cellulose to monosaccharides
[24, 98]. These microalgae can be used for ethanol production via various
hydrolysis strategies and fermentation processes [24, 93, 99] (Tables
7.1 and 7.6; Figure 7.10). Nahak et al. [105] reported that seaweed and
marine algae such as Eneteromorpha sp. contain 70% carbohydrate (dry
weight basis), which can be explored for bioethanol production [105]. The
reported microalgal species which are employed as a fermentable substrate
for bioethanol production are listed in Table 7.6. The microalgal cells
store carbohydrates in the outer layer of cell wall: pectin, agar, alginate;
inner layer of the cell wall: cellulose, hemicelluloses; and inside the cell:
starch [91]. These carbohydrates are hydrolyzed to fermentable sugars like
glucose for the bioethanol production via microbial fermentation [48, 72].
The procedure includes cultivation of microalgae followed by the recovery
of microalgal biomass from the medium, cell disruption for the release
of biomolecules, saccharification (hydrolysis), fermentation, and finally
separation by distillation [51, 91]. Chemical hydrolysis method facilitates
the solvent extraction of lipids, thereby recovering both fermentable sugars
and lipid from the microalgal biomass, where fermentable sugar can be
converted to ethanol and lipid into biodiesel, increasing the commercial
potential by two folds [106].
Algae-based fuel industries are still searching for innovative ways to
reduce the production cost via reducing the costs of systems infrastructure
and integration, algae biomass production process, harvesting, and dewa-
tering techniques, extraction, and fractionation, and finally biofuels conver-
sion process [107] (Figures 7.8–7.10).
Author Copy
Chlorella vulgaris 50.9 – 48.0 0.209 [24]
Chlorella vulgaris – – – 0.400 [103]
Dunaliella sp. – – – 0.011* [104]
*
Theoretical conversion.
FIGURE 7.9
Author Copy
Degradation products occurring during hydrolysis of lignocellulosic material [83].
Author Copy
FIGURE 7.10 Ethanol production from microalgae [51].
TABLE 7.7 List of Genetically Modified Microorganisms Used in Ethanol Production [110]
Microorganisms Strain Features
Yeast Candida shehatae NCL-3501 Co-ferment xylose and glucose
Apple Academic Press
Author Copy
xylose
Thermoanaerobacterium Improved ethanol yield, have
saccharolyticum ALK2 ability to ferment arabinose,
glucose, xylose, and mannose
Thermoanaerobacter mathranii Improved ethanol yield
BG1L1
Clostridium thermocellum
DSM1313 and YD01
Escherichia coli KO11 Ferment xylose and glucose
Escherichia coli FBR5 Ferment xylose and arabinose
Pichia stipites A Adapted at hydrolysate increased
Pichia stipitis NRRL Y-7124 concentration
7.5 PRETREATMENT
Author Copy
FIGURE 7.11 Ethanol biosynthesis pathway in Saccharomyces cerevisiae [111].
biological [126]. These methods differ from each other in context to mode of
action, reaction conditions and overall outcomes.
Physical method includes uncatalyzed steam explosion (SE), liquid
hot water (LHW) pretreatment, mechanical comminution and high energy
Apple Academic Press
radiation [124, 127]. The methods increase the surface area and pore volume,
decreases the degree of polymerization (DP) of cellulose and lignin [128].
By using uncatalyzed SE method, Grous et al. [129] achieved the enzymatic
digestibility of untreated poplar chips from 15% to 90% after treatment
[129]. Perez et al. [130] used LHW to pretreat wheat straw and obtained
maximum hemicellulose-derived sugar recovery of 53% and enzymatic
hydrolysis yield of 96% [130].
Author Copy
Although, the physical pretreatment method has some disadvantages like
energy consuming, ecologically unhealthy, and non-viable for commercial
process. Chemical pretreatments were initially exploited for the production
of high-quality paper products. The improvement in biodegradability of
cellulose by removing lignin and hemicellulose was the primary goal [124].
It includes techniques like alkali (e.g., NaOH, KOH, NH4OH and Ca(OH)2),
acid (H2SO4, HCl, HNO3), organic acids (fumaric and maleic acids) and
several cellulose solvents such as alkaline H2O2, ozone, glycerol, dioxane,
phenol, and ethylene glycol [124, 131, 132]. Millet et al. [133] reported an
increase in digestibility of NaOH-treated hardwood from 14% to 55% with a
decrease of lignin content from 24%–55% to 20% [133].
Physicochemical pretreatment is a combination of physical conditions
and chemicals. It includes SE, ammonia fiber explosion (AFEX), soaking
aqueous ammonia (SAA), ammonia recycling percolation (ARP), wet oxida-
tion and CO2 explosion. It increases the accessible surface area, decreases
cellulose crystallinity, and removes hemicellulose and lignin from the ligno-
cellulosic materials. The yields of enzymatic hydrolysis of AFEX-pretreated
newspaper (18–30% lignin) and aspen chips (25% lignin) were reported by
McMillan [23] as only 40% and below 50%, respectively [23].
Biological pretreatment is a friendly method used for lignin removal
[22]. It is carried out using microorganisms, particularly fungi, which are
white rot, brown rot, and soft rot fungi [134]. White and soft rot fungi
attack on both cellulose and lignin, whereas, brown rot fungi attack only
on cellulose. The biological pretreatment was studied by Hwang et al. [135]
using four different white-rot fungi for 30 days [135]. The team found that
glucose yield of pretreated wood by Trametes versicolor MrP 1 reached
45% by enzymatic hydrolysis while 35% solid was converted to glucose
during fungi incubation. Beside this, the lower hydrolysis rate and longer
incubation period as compared to the other pretreatment methods are some
Author Copy
inexpensive than the enzymatic method. However, the extreme conditions
like high temperature and pressure can degrade carbohydrate into furfural,
acetic acid, gypsum, vanillin, and aldehydes which are potential fermentation
inhibitors leading to yield reduction [91, 132].
The enzymatic hydrolysis of the carbohydrate, particularly starch, by
alpha-amylase (liquefaction) in a random manner generates oligosaccharides
with three or more ɑ-(1→4)-linked D-glucose units as an intermediate
product. Subsequently, starch saccharifying enzyme (amyloglucosidase) is
introduced to the liquefied starch and simple reducing sugars are generated
[83, 100]. This method has advantages over the chemical method such as
higher conversion yield, minimal byproduct formation, mild operating condi-
tion, and low energy input [138].
Author Copy
Two different reactors used for the hydrolysis and fermentation process
[142]. This has the advantage that both the process can be carried out under
different conditions like temperature, pH, and time [101, 102, 142]. Thus,
more substrate could be produced during the hydrolysis process by using the
In a single reactor, both hydrolysis and fermentation process take place [144].
Since, yeast will directly convert the glucose and cellobiose into bioethanol,
these compounds could not cause the inhibition [142]. SSF has advantages
Author Copy
over the SHF that it requires lower enzyme load, higher bioethanol yield,
shorter fermentation time, and reduced risk of contamination by external
microflora [145]. However, compromises in the operating conditions are
seen during the SSF process [146]. Additionally, it is difficult to recycle the
enzymes and yeast, which makes the process challenging for the scaling up
for commercialization purposes [147].
In the process, the same microorganisms carry out both saccharification and
fermentation of biomass in a single stage [148]. CBP is also known as direct
microbial conversion (DMC). CBP uses a single microorganism community
for cellulase production, cellulose hydrolysis, and fermentation in a single step,
has two major benefits: (i) avoid the cost of cellulase production, (ii) saves the
energy [149, 150]. Clostridium thermohydrosulfuricum, Thermoanaerobacter
ethanolicus, Thermoanaerobacter mathranii, Thermoanaerobacter brockii
strain, etc., are some thermophilic cellulolytic anaerobic bacteria that have
been studied for their potential as bioethanol producers [13].
Author Copy
factors which also influence the commercialization of ethanol production
including the availability of feedstock, cost of raw material and ethanol
recovery cost. The prices of feedstock depend on the location, seasons, local
state of the supply-demand conditions and transportation needed [152].
Studies are done which state that biomass feedstock made up 40% of the
ethanol production cost [139, 152]. The main challenge is to economize the
operating costs of biomass conversion processes that include pretreatment
and enzymatic hydrolysis, toward a full-scale commercialization. Extensive
studies are done with aim of developing a process which reduce bioconver-
sion time, use of cellulose during fermentation and enhance the total ethanol
yield [152, 153]. Another cost adding factor is the separation and recovery of
ethanol from the fermented liquor of lignocelluloses hydrolysates [136, 151].
Although microalgae are rich in carbohydrates and proteins, limited reports
are available on its use as a source for fermentation for ethanol production
[42]. Thus, despite being a fully functional industry, ethanol production
still has lacunae that need to be addressed to improve the efficiency of this
commercial endeavor and its economics.
7.8 CONCLUSION
With the sharp increase in population and vehicles, only bioethanol will not
be sufficient to match the energy demand. However, with each generation,
the bioprocess technology is undergoing massive changes, through which
we are slowly but contently increasing production potential. The changes
initiate from the nature of substrate to the microorganisms and the process
involved. Dealing with the current scenario of possible feedstock with their
advantages and disadvantages, it becomes necessary to evolve better and
sustainable methodologies for higher yield outputs. Feedstock generated
from forest and agricultural wastes, algal biomass and engineered organisms
could lead to the development of efficient technologies, thereby reducing the
ethanol price and environmental pollution, and also preserving the natural
resources. The fourth-generation biofuels are a new concept and could have
Apple Academic Press
ACKNOWLEDGMENTS
Author Copy
Center of Education” for research in basic sciences sanctioned under
CURIE program at the Department of Science and Technology.
CONFLICTS OF INTEREST
KEYWORDS
• bioethanol
• feedstock
• fermentation
• pretreatment
REFERENCES
1. Zhang, Y., (2013). The links between the price of oil and the value of US dollar. Int. J.
Energy Econ. Policy, 3, 341–351.
2. Lal, R., (2005). World crop residues production and implications of its use as a biofuel.
Environ. Int., 31, 575–584. https://doi.org/10.1016/j.envint.2004.09.005.
3. Rosales-Calderon, O., & Arantes, V., (2019). A review on commercial-scale high-value
products that can be produced alongside cellulosic ethanol. Biotechnol. Biofuels, 12, 240.
https://doi.org/10.1186/s13068‑019‑1529‑1.
4. Gaurav, N., Sivasankari, S., Kiran, G. S., Ninawe, A., & Selvin, J., (2017). Utilization of
bioresources for sustainable biofuels: A review. Renew. Sust. Energy Rev., 73, 205–214.
https://doi.org/ 10.1016/j.rser.2017.01.070.
5. Kremer, F. G., Jardim, J. L. F., & Maia, D. M., (1996). Effects of Alcohol Composition
on Gasohol Vehicle Emissions (No. 962094). SAE Technical Paper.
6. Demirbas, A., (2005). Bioethanol from cellulosic materials: A renewable motor fuel from
biomass. Energ. Sources, 27(4), 327–337. https://doi.org/10.1080/00908310390266643.
Apple Academic Press
7. Green, K. R., & Lowenbach, W. A., (2001). MTBE contamination: Environmental, legal,
and public policy challenges guest editorial. Environ. Forensics, 2(1), 3–6. https://doi.
org/10.1006/enfo.2001.0030.
8. Kar, Y., & Deveci, H., (2006). Importance of P-series fuels for flexible-fuel vehicles
(FFVs) and alternative fuels. Energ. Sources Part A, 28(10), 909–921. https://doi.org/
10.1080/00908310600718841.
9. Pickett, J., Anderson, D., Bowles, D., Bridgwater, T., Jarvis, P., Mortimer, N., Poliakoff,
M., & Woods, J., (2008). Sustainable Biofuels: Prospects and Challenges. The Royal
Society. UK. London.
Author Copy
10. Yao, C., Yang, X., Roy, R. R., Cheng, C., Tian, Z., & Li, Y., (2009). The effects of MTBE/
ethanol additives on toxic species concentration in gasoline flame. Energ. Fuel, 23(7),
3543–3548. https://doi.org/10.1021/ef900035q.
11. Minteer, S., (2006). Ethanol blends: E10 and E-diesel. Alcoholic Fuels, 125–136.
12. Nigam, P. S., & Singh, A., (2011). Production of liquid biofuels from renewable resources.
Prog. Energ. Combust., 37(1), 52–68. https://doi.org/10.1016/j.pecs.2010.01.003.
13. Vohra, M., Manwar, J., Manmode, R., Padgilwar, S., & Patil, S., (2014). Bioethanol
production: Feedstock and current technologies. J. Environ. Chem. Eng., 2, 573–584.
https://doi.org/10.1016/j.jece.2013.10.013.
14. Banerjee, S., Mudliar, S., Sen, R., Giri, B., Satpute, D., Chakrabarti, T., & Pandey, R.
A., (2010). Commercializing lignocellulosic bioethanol: Technology bottlenecks and
possible remedies. Biofuel. Bioprod. Bior., 4(1), 77–93. https://doi.org/10.1002/bbb.188.
15. Niphadkar, S., Bagade, P., & Ahmed, S., (2017). Bioethanol production: Insight into past,
present and future perspectives. Biofuels. http://dx.doi.org/10.1080/17597269.2017.133
4338.
16. Hunt, D. V., (1981). The Gasohol Handbook. New York, NY: Industrial Press.
17. Barros, S. (2019). “Brazil Biofuels Annual Report 2019.” United States Department of
Agriculture (USDA)–Foreign Agricultural Service (FAS). GAIN Report Number BR19029.
18. Ramos, J. L., & Duque, E., (2019). Twenty-first-century chemical odyssey: Fuels versus
commodities and cell factories versus chemical plants. Microb. Biotechnol., 12(2), 200–209.
https://doi.org/10.1111/1751-7915.13379.
19. Alonso-Gómez, L. A., & Bello-Pérez, L. A., (2018). Four generations of raw materials
used for ethanol production: Challenges and opportunities. Agrociencia (Montecillo),
52(7), 967–990.
20. Dalena, F., Senatore, A., Iulianelli, A., Di Paola, L., Basile, M., & Basile, A., (2019).
Ethanol from biomass: Future and perspectives. In: Ethanol (pp. 25–59). Elsevier. https://
doi.org/10.1016/B978-0-12-811458-2.00002-X.
21. Alia, K. B., Rasul, I., Azeem, F., Hussain, S., Siddique, M. H., Muzammil, S., Riaz, M.,
et al., (2019). Microbial production of ethanol. In: Microbial Fuel Cells: Materials and
Applications (pp. 307–334). Materials Research Forum LLC., Pennsylvania. https://doi.
org/10.21741/9781644900116-12.
22. Ahorsu, R., Medina, F., & Constantí, M., (2018). Significance and challenges of biomass
as a suitable feedstock for bioenergy and biochemical production: A review. Energies,
11(12), 3366. https://doi.org/10.3390/en11123366.
23. McMillan, J. D., (1993). Xylose Fermentation to Ethanol: A Review (No. NREL/
TP-421-4944). National Renewable Energy Lab., Golden, CO (United States).
24. Ho, D. P., Ngo, H. H., & Guo, W., (2014). A mini-review on renewable sources for biofuel.
Bioresour. Technol., 169, 742–749. https://doi.org/10.1016/j.biortech.2014.07.022
Apple Academic Press
25. Sindhu, R., Gnansounou, E., Binod, P., & Pandey, A., (2016). Bioconversion of sugarcane
crop residue for value-added products: An overview. Renew. Energ., 98, 203–215. https://
doi.org/10.1016/j.renene.2016.02.057.
26. Kim, S., & Dale, B. E., (2004). Global potential bioethanol production from wasted
crops and crop residues. Biomass Bioenerg., 26(4), 361–375. https://doi.org/10.1016/j.
biombioe.2003.08.002.
27. Zabed, H., Sahu, J. N., Boyce, A. N., & Faruq, G., (2016). Fuel ethanol production
from lignocellulosic biomass: An overview on feedstocks and technological approaches.
Renew. Sust. Energ. Rev., 66, 751–774. https://doi.org/10.1016/j.rser.2016.08.038.
Author Copy
28. Ramírez, M. B., Ferrari, M. D., & Lareo, C., (2016). Fuel ethanol production from
commercial grain sorghum cultivars with different tannin content. J. Cereal Sci., 69,
125–131. https://doi.org/10.1016/j.jcs.2016.02.019.
29. Murphy, J. D., & Power, N. M., (2008). How can we improve the energy balance of
ethanol production from wheat?. Fuel, 87(10, 11), 1799–1806. https://doi.org/10.1016/j.
fuel.2007.12.011.
30. Pervez, S., Aman, A., Iqbal, S., Siddiqui, N. N., & Qader, S. A. U., (2014). Saccharification
and liquefaction of cassava starch: An alternative source for the production of bioethanol
using amylolytic enzymes by double fermentation process. BMC Biotechnol., 14(1), 49.
https://doi.org/10.1186/1472-6750-14-49.
31. Ziska, L. H., Runion, G. B., Tomecek, M., Prior, S. A., Torbet, H. A., & Sicher, R.,
(2009). An evaluation of cassava, sweet potato and field corn as potential carbohydrate
sources for bioethanol production in Alabama and Maryland. Biomass Bioenerg.,
33(11), 1503–1508. https://doi.org/10.1016/j.biombioe.2009.07.014.
32. Khan, R. A., Nawaz, A., Ahmed, M., Khan, M. R., Azam, F. D., Ullah, S., Sadullah, F.,
et al., (2012). Production of bioethanol through enzymatic hydrolysis of potato. Afr. J.
Biotechnol., 11(25), 6739–6743. http://dx.doi.org/10.5897/AJB11.2791.
33. George, N. A., Pecota, K. V., Bowen, B. D., Schultheis, J. R., & Yencho, G. C., (2011). Root
piece planting in sweet potato—A synthesis of previous research and directions for the
future. Horttechnology, 21(6), 703–711. https://doi.org/10.21273/HORTTECH.21.6.703.
34. Duvernay, W. H., Chinn, M. S., & Yencho, G. C., (2013). Hydrolysis and fermentation
of sweet potatoes for production of fermentable sugars and ethanol. Ind. Crops Prod.,
42, 527–537. https://doi.org/10.1016/j.indcrop.2012.06.028.
35. Lewandowski, I., Scurlock, J. M., Lindvall, E., & Christou, M., (2003). The development
and current status of perennial rhizomatous grasses as energy crops in the US and Europe.
Biomass Bioenerg., 25(4), 335–361.https://doi.org/10.1016/S0961-9534(03)00030-8.
36. Rengsirikul, K., Ishii, Y., Kangvansaichol, K., Sripichitt, P., Punsuvon, V., Vaithanomsat,
P., Nakamanee, G., & Tudsri, S., (2013). Biomass yield, chemical composition
and potential ethanol yields of 8 cultivars of Napier grass (Pennisetum purpureum
Schumach.) harvested 3-Monthly in central Thailand. J. Sust. Bioenergy Syst., 3(2),
107–112. https://doi.org/10.4236/jsbs.2013.32015.
37. Kumar, A., Singh, L. K., & Ghosh, S., (2009). Bioconversion of lignocellulosic fraction
of water-hyacinth (Eichhornia crassipes) hemicellulose acid hydrolysate to ethanol
by Pichia stipitis. Bioresour. Technol., 100(13), 3293–3297. https://doi.org/10.1016/j.
biortech.2009.02.023.
38. Aswathy, U. S., Sukumaran, R. K., Devi, G. L., Rajasree, K. P., Singhania, R. R., &
Pandey, A., (2010). Bio-ethanol from water hyacinth biomass: An evaluation of enzymatic
saccharification strategy. Bioresour. Technol., 101(3), 925–930. https://doi.org/10.1016/j.
biortech.2009.08.019.
Apple Academic Press
39. Zhu, J. Y., & Pan, X. J., (2010). Woody biomass pretreatment for cellulosic ethanol
production: Technology and energy consumption evaluation. Bioresour. Technol., 101(13),
4992–5002. https://doi.org/10.1016/j.biortech.2009.11.007.
40. Gonzalez, R. W., Treasure, T., Phillips, R. B., Jameel, H., & Saloni, D., (2011). Economics
of cellulosic ethanol: Green liquor pretreatment for softwood and hardwood, greenfield
and repurpose scenarios. Bioresources, 6(3), 2551–2567.
41. Zhu, J. Y., Luo, X., Tian, S., Gleisner, R., Negrone, J., & Horn, E., (2011). Efficient
ethanol production from beetle-killed lodgepole pine using SPORL technology and
Saccharomyces cerevisiae without detoxification. Tappi J., 10(5), 9–18.
Author Copy
42. Singh, J., & Gu, S., (2010). Commercialization potential of microalgae for biofuels produc-
tion. Renew. Sust. Energ. Rev., 14(9), 2596–2610. https://doi.org/10.1016/j.rser.2010.06.014.
43. Limayem, A., & Ricke, S. C., (2012). Lignocellulosic biomass for bioethanol production:
Current perspectives, potential issues and future prospects. Prog. Energ. Comb., 38(4),
449–467. https://doi.org/10.1016/j.pecs.2012.03.002.
44. Kreith, F., & Krumdieck, S., (2013). Principles of Sustainable Energy Systems. CRC Press.
45. Buah, W. K., Cunliffe, A. M., & Williams, P. T., (2007). Characterization of products
from the pyrolysis of municipal solid waste. Process Saf. Environ., 85(5), 450–457.
https://doi.org/10.1205/psep07024.
46. Prasad, S., Singh, A., & Joshi, H. C., (2007). Ethanol as an alternative fuel from agricultural,
industrial and urban residues. Resour. Conserv. Recy., 50(1), 1–39. https://doi.org/10.1016/j.
resconrec.2006.05.007.
47. Li, S., Zhang, X., & Andresen, J. M., (2012). Production of fermentable sugars from
enzymatic hydrolysis of pretreated municipal solid waste after autoclave process. Fuel,
92(1), 84–88. https://doi.org/10.1016/j.fuel.2011.07.012.
48. Lam, M. K., & Lee, K. T., (2015). Bioethanol production from microalgae. In: Handbook of
Marine Microalgae (pp. 197–208). https://doi.org/10.1016/B978-0-12-800776-1.00012-1.
49. Nguyen, M. A., & Hoang, A. L., (2016). A Review on Microalgae and Cyanobacteria in
Biofuel Production. USTH: Hanoi, Vietnam.
50. Duden, A. S., Verweij, P. A., Faaij, A. P. C., Baisero, D., Rondinini, C., & Van, D. H. F.,
(2020). Biodiversity impacts of increased ethanol production in Brazil. Land, 9(1), 12.
https://doi.org/10.3390/land9010012.
51. Mussatto, S. I., Dragone, G., Guimarães, P. M. R., Paulo, J., Silva, A., Carneiro, I.
M., Roberto, I. C., et al., (2010). Technological trends, global market, and challenges
of bio-ethanol production. Biotechnol. Adv., 28(6), 817–830. https://doi.org/10.1016/j.
biotechadv.2010.07.001.
52. Cheng, J. J., & Timilsina, G. R., (2011). Status and barrier of advanced biofuel technologies:
A review. Renew. Energ., 36, 3541–3549. https://doi.org/10.1016/j.renene.2011.04.031.
53. Erdal, G., Esengün, K., Erdal, H., & Gündüz, O., (2007). Energy use and economical
analysis of sugar beet production in Tokat province of Turkey. Energy, 32(1), 35–41.
https://doi.org/10.1016/j.energy.2006.01.007.
54. Cosyn, S., Woude, K. V. D., Sauvenier, X., & Evrard, J. N., (2011). Sugar beet: A
complement to sugar cane for sugar and ethanol production in tropical and subtropical
areas. Int. Sugar J., 113(1346), 120–123.
55. Duraisam, R., Salelgn, K., & Berekete, A. K., (2017). Production of beet sugar and bioethanol
from sugar beet and it bagasse: A Review. Int. J. Eng. Trends Technol., 43, 222–233.
56. De La Piscina, P. R., & Homs, N., (2008). Use of biofuels to produce hydrogen (reformation
processes). Chem. Soc. Rev., 37(11), 2459–2467. https://doi.org/10.1039/B712181B.
Apple Academic Press
57. Grahovac, J. A., Dodić, J. M., Dodić, S. N., Popov, S. D., Vučurović, D. G., & Jokić,
A. I., (2012). Future trends of bioethanol co-production in Serbian sugar plants. Renew.
Sust. Energ. Rev., 16(5), 3270–3274. https://doi.org/10.1016/j.rser.2012.02.040.
58. Dien, B. S., Cotta, M. A., & Jeffries, T. W., (2003). Bacteria engineered for fuel ethanol
production: Current status. Appl. Microbiol. Biotechnol., 63(3), 258–266. https://doi.
org/10.1007/s00253-003-1444-y.
59. Jeffries, T. W., & Jin, Y. S., (2004). Metabolic engineering for improved fermentation of
pentoses by yeasts. Appl. Microbiol. Biotechnol., 63(5), 495–509. https://doi.org/10.1007/
s00253-003-1450-0.
Author Copy
60. Jobling, S., (2004). Improving starch for food and industrial applications. Curr. Opin.
Plant Biol., 7(2), 210–218. https://doi.org/10.1016/j.pbi.2003.12.001.
61. Solomon, B. D., Barnes, J. R., & Halvorsen, K. E., (2007). Grain and cellulosic ethanol:
History, economics, and energy policy. Biomass Bioenerg., 31(6), 416–425. https://doi.
org/10.1016/j.biombioe.2007.01.023.
62. Khanal, S. K., (2008). anaerobic biotechnology for bioenergy production: Principles and
applications. In: Conference Proceedings (pp. 43–63).
63. Bothast, R. J., & Schlicher, M. A., (2005). Biotechnological processes for conversion of
corn into ethanol. Appl. Microbiol. Biotechnol., 67(1), 19–25. https://doi.org/10.1007/
s00253-004-1819-8.
64. Erickson, G. E., Klopfenstein, T. J., Adams, D. C., & Rasby, R. J., (2005). General
overview of feeding corn milling co-products to beef cattle. Corn Processing Co-products
Manual, 3–12.
65. Kim, Y., Mosier, N., & Ladisch, M. R., (2008). Process simulation of modified dry grind
ethanol plant with recycle of pretreated and enzymatically hydrolyzed distillers’ grains.
Bioresour. Technol., 99(12), 5177–5192. https://doi.org/10.1016/j.biortech.2007.09.035.
66. Ziolkowska, J. R., (2020). Biofuels technologies: An overview of feedstocks, processes,
and technologies. In: Biofuels for a More Sustainable Future (pp. 1–19). Elsevier.
67. Baffes, J., (2013). A framework for analyzing the interplay among food, fuels, and
biofuels. Global Food Security, 2(2), 110–116. https://doi.org/10.1016/j.gfs.2013.04.003.
68. Filip, O., Janda, K., Kristoufek, L., & Zilberman, D., (2017). Food versus fuel: An updated
and expanded evidence. Energy Econ. https://doi.org/10.1016/j.eneco.2017.10.033.
69. Tomei, J., & Helliwell, R., (2016). Food versus fuel? Going beyond biofuels. Land Use
Policy, 56, 320–326. https://doi.org/10.1016/j.landusepol.2015.11.015.
70. Isikgor, F. H., & Becer, C. R., (2015). Lignocellulosic biomass: A sustainable platform for
the production of bio-based chemicals and polymers. Polym. Chem., 6(25), 4497–4559.
https://doi.org/ 10.1039/C5PY00263J.
71. Balat, M., & Balat, H., (2009). Recent trends in global production and utilization of bioethanol
fuel. Appl. Energ., 86, 2273–2282. https://doi.org/10.1016/j.apenergy. 2009.03.015.
72. Harun, R., Singh, M., Forde, G. M., & Danquah, M. K., (2010). Bioprocess engineering
of microalgae to produce a variety of consumer products. Renew. Sust. Energ. Rev., 14,
1037–1047. https://doi.org/10.1016/j.rser.2009.11.004.
73. Naik, S. N., Goud, V. V., Rout, P. K., & Dalai, A. K., (2010). Production of first and
second-generation biofuels: A comprehensive review. Renew. Sust. Energ. Rev., 14(2),
578–597. https://doi.org/10.1016/j.rser.2009.10.003.
74. Dragone, G., Fernandes, B. D., Vicente, A. A., & Teixeira, J. A., (2010). Third-generation
biofuels from microalgae. Curr. Res. Technol. Edu. Topics. Appl. Microbiol. Microbial.
Biotechnol., 2, 1355–1366.
75. Elshahed, M. S., (2010). Microbiological aspects of biofuel production: Current status and
Apple Academic Press
Author Copy
fermentation using a cellulase of paper sludge origin and thermotolerant Saccharomyces
cerevisiae TJ14. Biotechnol. Biofuels, 4, 35. https://doi.org/10.1186/1754-6834-4-35.
79. Robak, K., & Balcerek, M., (2018). Review of second-generation bioethanol production
from residual biomass. Food Technol. Biotechnol., 56(2), 174–187. https://doi.org/10.17113/
ftb.56.02.18.5428.
80. Alvira, P., Tomás-Pejó, E., Ballesteros, M., & Negro, M. J., (2010). Pretreatment technolo-
gies for an efficient bioethanol production process based on enzymatic hydrolysis: A review.
Bioresour. Technol., 101(13), 4851–4861. https://doi.org/10.1016/j.biortech.2009.11.093.
81. Cardona, C. A., & Sánchez, Ó. J., (2007). Fuel ethanol production: Process design
trends and integration opportunities. Bioresour. Technol., 98(12), 2415–2457. https://
doi.org/10.1016/j.biortech.2007.01.002.
82. Tiwari, G., Sharma, S., & Prasad, R., (2015). Bioethanol production: Future prospects
from non-traditional sources in India. Int. J. Res. BioSci., 4(4), 1–15.
83. Demirbas, A., (2008). Products from lignocellulosic materials via degradation processes.
Energ. Sources Part A, 30(1), 27–37. https://doi.org/10.1080/00908310600626705.
84. Czekała, W., Bartnikowska, S., Dach, J., Janczak, D., & Mazurkiewicz, J., (2018). The
energy value and economic efficiency of solid biofuels produced from digestate and
sawdust. Energy, 159, 1118–1122. https://doi.org/10.1016/j.energy.2018.06.090.
85. Ge, Y., & Li, L., (2018). System-level energy consumption modeling and optimization for
cellulosic biofuel production. Appl. Energy, 226(15), 935–946. https://doi.org/10.1016/j.
apenergy.2018.06.020.
86. Ziolkowska, J. R., (2014). Prospective technologies, feedstocks and market innovations
for ethanol and biodiesel production in the US. Biotechnol. Rep., 4, 94–98. https://doi.
org/10.1016/j.btre.2014.09.001.
87. Bhatia, S. K., Kim, S. H., Yoon, J. J., & Yang, Y. H., (2017). Current status and strategies
for second-generation biofuel production using microbial systems. Energy Convers.
Manag., 148(15), 1142–1156. https://doi.org/10.1016/j.enconman.2017.06.073.
88. Suganya, T., Varman, M., Masjuki, H. H., & Renganathan, S., (2016). Macroalgae
and microalgae as a potential source for commercial applications along with biofuels
production: A biorefinery approach. Renew. Sust. Energ. Rev., 55, 909–941. http://dx.doi.
org/10.1016/j.rser.2015.11.026.
89. Kim, M., & Day, D. F., (2011). Composition of sugar cane, energy cane, and sweet
sorghum suitable for ethanol production at Louisiana sugar mills. J. Ind. Microbiol.
Biotechnol., 38(7), 803–807.https://doi.org/10.1007/s10295-010-0812-8.
90. Özçimen, D., & İnan, B., (2015). An overview of bioethanol production from algae.
Biofuels-Status and Perspective, 141–162.
91. Chen, C. Y., Zhao, X. Q., Yen, H. W., Ho, S. H., Cheng, C. L., Lee, D. J., Bai, F. W., &
Chang, J. S., (2013). Microalgae-based carbohydrates for biofuel production. Biochem.
Apple Academic Press
Author Copy
6, 1381–1389.
95. Thatoi, H., Dash, P. K., Mohapatra, S., & Swain, M. R., (2016). Bioethanol production
from tuber crops using fermentation technology: A review. Int. J. Sust. Energ., 35(5),
443–468. https://doi.org/10.1080/14786451.2014.918616.
96. Song, D., Fu, J., & Shi, D., (2008). Exploitation of oil-bearing microalgae for biodiesel.
Chin. J. Biotechnol., 24(3), 341–348.https://doi.org/10.1016/S1872-2075(08)60016-3
97. Tredici, M. R., (2010). Photobiology of microalgae mass cultures: Understanding
the tools for the next green revolution. Biofuels, 1, 143–162. https://doi.org/10.4155/
bfs.09.10.
98. Harun, R., Yip, J. W., Thiruvenkadam, S., Ghani, W. A., Cherrington, T., & Danquah,
M. K., (2014). Algal biomass conversion to bioethanol-a step-by-step assessment.
Biotechnol. J., 9(1), 73–86. https://doi.org/10.1002/biot.201200353.
99. Varaprasad, D., Narasimham, D., Paramesh, K., Sudha, N. R., Himabindu, Y., Keerthi,
K. M., Nazaneen, P. S., & Chandrasekhar, T., (2019). Improvement of ethanol production
using green alga Chlorococcum minutum. Env. Technol., 1–9. https://doi.org/10.1080/0
9593330.2019.1669719.
100. Choi, S. P., Nguyen, M. T., & Sim, S. J., (2010). Enzymatic pretreatment of Chlamydomonas
reinhardtii biomass for ethanol production. Bioresour. Technol., 101(14), 5330–5336.
https://doi.org/10.1016/j.biortech.2010.02.026.
101. Harun, R., & Danquah, M. K., (2011). Influence of acid pretreatment on microalgal
biomass for bioethanol production. Process Biochem., 46(1), 304–309. https://doi.
org/10.1016/j.procbio.2010.08.027.
102. Harun, R., Jason, W. S. Y., Cherrington, T., & Danquah, M. K., (2011). Exploring
alkaline pretreatment of microalgal biomass for bioethanol production. Appl. Energ.,
88(10), 3464–3467. https://doi.org/10.1016/j.apenergy.2010.10.048.
103. Lee, S., Oh, Y., Kim, D., Kwon, D., Lee, C., & Lee, J., (2011). Converting carbohydrates
extracted from marine algae into ethanol using various ethanolic Escherichia coli strains.
Appl. Biochem. Biotechnol., 164(6), 878–888. https://doi.org/10.1007/s12010-011-9181-7.
104. Shirai, F., Kunii, K., Sato, C., Teramoto, Y., Mizuki, E., Murao, S., & Nakayama, S.,
(1998). Cultivation of microalgae in the solution from the desalting process of soy sauce
waste treatment and utilization of the algal biomass for ethanol fermentation. World.
J. Microbiol. Biotechnol., 14(6), 839–842. https://doi.org/10.1023/A:1008860705434.
105. Nahak, S., Nahak, G., Pradhan, I., & Sahu, R. K., (2011). Bioethanol from marine algae:
A solution to global warming problem. J. Appl. Environ. Biol. Sci., 1(4), 74–80.
106. Silva, C. E. D. F., & Bertucco, A., (2019). Bioethanol from microalgal biomass: A promising
approach in biorefinery. Braz. Arch. Biol. Technol., 62. http://dx.doi.org/10.1590/1678-
4324-2019160816.
107. US DOE, (2010). National Algal Biofuels Technology Roadmap. DOE, Washington, DC.
Apple Academic Press
108. Dutta, K., Daverey, A., & Lin, J. G., (2014). Evolution retrospective for alternative fuels:
First to fourth generation. Renew. Energy, 69, 114–122. https://doi.org/10.1016/j.renene.
2014.02.044.
109. Brethauer, S., & Studer, M. H., (2014). Consolidated bioprocessing of lignocellulose by
a microbial consortium. Energy Environ. Sci., 7(4), 1446–53. https://doi.org/10.1039/
c3ee41753k.
110. Aditiya, H. B., Mahlia, T. M. I., Chong, W. T., Nur, H., & Sebayang, A. H., (2016).
Second-generation bioethanol production: A critical review. Renew. Sustain. Energy
Rev., 66, 631–653. https://doi.org/10.1016/j.rser.2016.07.015.
Author Copy
111. Macedo, N., & Brigham, C. J., (2014). From beverages to biofuels: The journeys of
ethanol-producing microorganisms. Int. J. Biotechnol. Wellness Ind., 3(3), 79–87. http://
dx.doi.org/10.6000/1927-3037.2014.03.03.1.
112. Xia, J., Yang, Y., Liu, C. G., Yang, S., & Bai, F. W., (2019). Engineering Zymomonas
mobilis for robust cellulosic ethanol production. Trends Biotechnol., (2019). https://doi.
org/10.1016/j.tibtech.2019.02.002.
113. Karsch, T., Stahl, U., & Esser, K., (1983). Ethanol production by Zymomonas and Saccharo-
myces, advantages and disadvantages. Eur. J. Appl. Microbiol. Biotechnol., 18(6), 387–391.
https://doi.org/10.1007/BF00504750.
114. Zhang, M., Eddy, C., Deanda, K., Finkelstein, M., & Picataggio, S., (1995). Metabolic
engineering of a pentose metabolism pathway in ethanologenic Zymomonas mobilis.
Science, 267(5195), 240–243. https://doi.org/10.1126/science.267.5195.240.
115. Quevedo-Hidalgo, B., Monsalve-Marín, F., Narváez-Rincón, P. C., Pedroza-Rodríguez, A.
M., & Velásquez-Lozano, M. E., (2013). Ethanol production by Saccharomyces cerevisiae
using lignocellulosic hydrolysate from Chrysanthemum waste degradation. World. J.
Microb. Biotechnol., 29(3), 459–466. https://doi.org/10.1007/s11274-012-1199-7.
116. Diaz, C. E., Sierra, Y. K., & Hernández, J. A., (2018). Determination of the percentage
of ethanol produced by Saccharomyces cerevisiae from semi-purified glycerin. In:
Journal of Physics: Conference Series (Vol. 1126, No. 1, p. 012008). https://doi.org/
10.1088/1742-6596/1126/1/012008.
117. Todhanakasem, T., Salangsing, O. L., Koomphongse, P., Kanokratana, P., & Champreda,
V., (2019). Zymomonas mobilis biofilm reactor for ethanol production using rice straw
hydrolysate under continuous and repeated batch processes. Front. Microbiol., 10, 1777.
https://doi.org/10.3389/fmicb.2019.01777.
118. Horn, S. J., Aasen, I. M., & Østgaard, K., (2000). Ethanol production from seaweed extract.
J. Ind. Microbiol. Biotechnol., 25(5), 249–254. https://doi.org/10.1038/sj.jim.7000065.
119. Harun, R., Danquah, M. K., & Forde, G. M., (2010). Microalgal biomass as a fermentation
feedstock for bioethanol production. J. Chem. Technol. Biotechnol., 85(2), 199–203.
https://doi.org/10.1002/jctb.2287.
120. Adams, J. M. M., Toop, T. A., Donnison, I. S., & Gallagher, J. A., (2011). Seasonal variation
in Laminaria digitata and its impact on biochemical conversion routes to biofuels.
Bioresour. Technol., 102(21), 9976–9984. https://doi.org/10.1016/j.biortech.2011.08.032.
121. Rastogi, M., & Shrivastava, S., (2017). Recent advances in second-generation bioethanol
production: An insight to pretreatment, saccharification and fermentation processes.
Renew. Sust. Energ. Rev., 80, 330–340. https://doi.org/10.1016/j.rser.2017.05.225.
122. Anisha, G. S., & John, R. P., (2015). Microalgae as an alternative feedstock for green
biofuel technology. Environ. Res. J., 9(2), 223–240.
123. Simas-Rodrigues, C., Villela, H. D., Martins, A. P., Marques, L. G., Colepicolo, P., &
Tonon, A. P., (2015). Microalgae for economic applications: Advantages and perspectives
Apple Academic Press
Author Copy
127. Taherzadeh, M. J., & Karimi, K., (2008). Pretreatment of lignocellulosic wastes to
improve ethanol and biogas production: A review. Int. J. Mol. Sci., 9(9), 1621–1651.
https://doi.org/10.3390/ijms9091621.
128. Szczodrak, J., & Fiedurek, J., (1996). Technology for conversion of lignocellulosic biomass
to ethanol. Biomass Bioenerg., 10(5, 6), 367–375. https://doi.org/10.1016/0961-9534
(95)00114-X.
129. Grous, W. R., Converse, A. O., & Grethlein, H. E., (1986). Effect of steam explosion
pretreatment on pore size and enzymatic hydrolysis of poplar. Enzyme Microb. Technol.,
8(5), 274–280. https://doi.org/10.1016/0141-0229(86)90021-9.
130. Pérez, J. A., González, A., Oliva, J. M., Ballesteros, I., & Manzanares, P., (2007). Effect
of process variables on liquid hot water pretreatment of wheat straw for bioconversion to
fuel-ethanol in a batch reactor. J. Chem. Technol. Biotechnol., 82(10), 929–938. https://
doi.org/10.1002/jctb.1765.
131. Sun, Y., & Cheng, J., (2002). Hydrolysis of lignocellulosic materials for ethanol production: A
review. Bioresour. Technol., 83(1), 1–11. https://doi.org/10.1016/S0960-8524 (01)00212-7.
132. Bensah, E. C., & Mensah, M., (2013). Chemical pretreatment methods for the production
of cellulosic ethanol: Technologies and innovations. Int. J. Chem. Eng. https://doi.org/
10.1155/2013/719607.
133. Millett, M. A., Baker, A. J., & Satter, L. D., (1976). Physical and chemical pretreatments
for enhancing cellulose saccharification. Biotechnol. Bioeng. Symp., 6, 125–153.
134. Sarkar, N., Ghosh, S. K., Bannerjee, S., & Aikat, K., (2012). Bioethanol production from
agricultural wastes: An overview. Renew. Energ., 37(1), 19–27. https://doi.org/10.1016/j.
renene.2011.06.045.
135. Hwang, S. S., Lee, S. J., Kim, H. K., Ka, J. O., Kim, K. J., & Song, H. G., (2008).
Biodegradation and saccharification of wood chips of Pinus strobus and Liriodendron
tulipifera by white-rot fungi. J. Microbiol. Biotechnol., 18, 1819–1825. https://doi.org/
10.4014/jmb.0800.231.
136. Knauf, M., & Moniruzzaman, M., (2004). Lignocellulosic biomass processing: A
perspective. Int. Sugar J., 106(1263), 147–150.
137. Miranda, J. R., Passarinho, P. C., & Gouveia, L., (2012). Pretreatment optimization of
Scenedesmus obliquus microalga for bioethanol production. Bioresour. Technol., 104,
342–348. https://doi.org/10.1016/j.biortech.2011.10.059.
138. Taherzadeh, M. J., & Karimi, K., (2007). Enzyme-based hydrolysis processes for ethanol
from lignocellulosic materials: A review. BioResources, 2(4), 707–738.
139. Hamelinck, C. N., Van, H. G., & Faaij, A. P., (2005). Ethanol from lignocellulosic biomass:
Techno-economic performance in short-, middle-and long-term. Biomass Bioenerg., 28(4),
384–410. https://doi.org/10.1016/j.biombioe.2004.09.002.
140. Talebnia, F., Karakashev, D., & Angelidaki, I., (2010). Production of bioethanol from
Apple Academic Press
Author Copy
144. Hahn-Hägerdal, B., Galbe, M., Gorwa-Grauslund, M. F., Lidén, G., & Zacchi, G.,
(2006). Bio-ethanol-the fuel of tomorrow from the residues of today. Trends Biotechnol.,
24(12), 549–556. https://doi.org/10.1016/j.tibtech.2006.10.004.
145. Lin, Y., & Tanaka, S., (2006). Ethanol fermentation from biomass resources: Current state
and prospects. Appl. Microbiol. Biotechnol., 69(6), 627–642. https://doi.org/10.1007/
s00253-005-0229-x.
146. Gupta, R. B., & Demirbas, A., (2010). Gasoline, Diesel, and Ethanol Biofuels from
Grasses and Plants. Cambridge University Press.
147. Olofsson, K., Bertilsson, M., & Lidén, G., (2008). A short review on SSF–an interesting
process option for ethanol production from lignocellulosic feedstocks. Biotechnol.
Biofuels, 1(1), 7. https://doi.org/10.1186/1754-6834-1-7.
148. Carere, C., Sparling, R., Cicek, N., & Levin, D., (2008). Third-generation biofuels via
direct cellulose fermentation. Int. J. Mol. Sci., 9(7), 1342–1360. https://doi.org/10.3390/
ijms9071342.
149. Lynd, L. R., Weimer, P. J., Van, Z. W. H., & Pretorius, I. S., (2002). Microbial cellulose
utilization: Fundamentals and biotechnology. Microbiol. Mol. Biol. Rev., 66(3), 506–577.
https://doi.org/10.1128/MMBR.66.3.506-577.2002.
150. Lynd, L. R., Van, Z. W. H., McBride, J. E., & Laser, M., (2005). Consolidated
bioprocessing of cellulosic biomass: An update. Curr. Opin. Biotechnol., 16(5), 577–583.
https://doi.org/10.1016/j.copbio.2005.08.009.
151. Sanchez, O. J., & Cardona, C. A., (2008). Trends in biotechnological production of fuel
ethanol from different feedstocks. Bioresour. Technol., 99(13), 5270–5295. https://doi.
org/10.1016/j.biortech.2007.11.013.
152. Chandel, A. K., Singh, O. V., Chandrasekhar, G., Rao, L. V., & Narasu, M. L., (2010).
Key drivers influencing the commercialization of ethanol-based biorefineries. J.
Commer. Biotechnol., 16(3), 239–257. https://doi.org/10.1057/jcb.2010.5.
153. Gray, K. A., Zhao, L., & Emptage, M., (2006). Bioethanol. Curr. Opin. Chem. Biol.,
10(2), 141–146.
Author Copy
Overijssel, The Netherland, E-mail: a.d.perezavila@utwente.nl
Department of Chemical Engineering, National University of Colombia,
2
ABSTRACT
8.1 INTRODUCTION
Author Copy
the thermodynamic limitations in distillation, the use of entrainer/solvents,
and the need of several number of columns in the distillation trains make
these technologies not very efficient to recover ethanol, especially to use it
as fuel alcohol.
In addition, adsorption has been successfully applied in several biorefineries
to recover ethanol from aqueous solutions. Extractive or azeotropic distillation
columns can be replaced by adsorption to overcome the azeotrope (ethanol
dehydration). For instance, the use of molecular sieves (from potassium
aluminosilicates) has been applied in industrial cases, achieving to replace
the azeotropic distillation [1]. Although adsorption with molecular sieves
can associate high capital costs, it is a technology with a high separation
performance with a lower operation cost as compared to traditional distillation
schemes.
On the other hand, pressure-swing distillation has been extensively studied
to separate ethanol by reducing the operation pressure inside a column in the
separation train. At vacuum condition, the separation by distillation leads
to the disappearance of the azeotrope. However, it is necessary the use of
columns composed of a large number of plates, and high reflux ratios to
achieve an ethanol with a high purity. Also, high capital and operation costs
can be associated to maintain the vacuum condition for a large volume [2].
Moreover, several efforts have been made in order to propose separation
technologies with a higher performance than those previously mentioned.
For instance, the use of non-conventional and non-volatile solvents to
increase the ethanol volatility, and consequently, to overcome the azeotropic
condition. In this chapter, ED using dissolved salts, ionic liquid and hyper-
branched polymers (HyPols) are presented.
Also, ethanol recovery through membrane technology and its integration
to distillation is described. Pervaporation (PV) has been demonstrated to
be able to separate ethanol without thermodynamic limitations, which is a
Author Copy
fermentation as alternative to increase the fermentation performance.
Author Copy
FIGURE 8.1 Distillation process to obtain hydrated ethanol from sugarcane molasses with
thermal integration in the column condenser.
Source: Adapted from: Dias et al. [7].
usually applied (Figure 8.1). The wine is used as a cooling service stream in
the condenser C2 at the top of column B, producing a partial condensation
of the hydrated ethanol, followed by a total condensation in condenser C3
with water. The wine leaves C2 at 60°C. The heated wine is used to reduce
the stillage temperature as well. The wine achieves a temperature of 90°C,
before being fed to the first distillation column.
The produced ethanol in biorefineries is mainly used as fuel alcohol.
Author Copy
Therefore, hydrated ethanol is not still applicable for blending in gasoline in
flex-fuel engines, which are the most common technology using fuel alcohol.
Azeotropic ethanol must be dehydrated. At least an ethanol composition of
99.3 wt.% or higher must be produced to be added in gasoline [8]. Since it is
not possible to overcome the azeotropic composition by conventional distil-
lation, azeotropic distillation (AD) is used to produce anhydrous ethanol.
AD is an economical technology widely applied in Brazilian distilleries [9].
Figure 8.2 is presented the flow sheet of a HTAD for the ethanol dehy-
dration. The hydrated ethanol and the entrainer are mixed and fed to the
azeotropic distillation column (1), where anhydrous alcohol is produced
at the bottom of the same column. The product at the top of the column
Apple Academic Press
Author Copy
again.
FIGURE 8.2 General scheme of azeotropic distillation for the separation of ethanol-water
mixtures: (1) azeotropic distillation column; (2) decanter; and (3) distillation column for the
entrainer recovery [12].
Author Copy
They are identified according to their nominal size of the internal pores
whose diameter is generally measured in angstroms. Molecular sieves are
materials that are characterized by their excellent capacity to retain on their
surface defined types of chemical species. These species are usually solvents
(water most of the cases), which are desired to be removed from a mixture to
obtain a final product with its given specifications [15].
Adsorption has been applied as an alternative process for ethanol
dehydration. Recently, new distilleries plants have opted to use adsorption
rather than AD to produce anhydrous ethanol. Adsorption on molecular
sieves is an established technology for the ethanol dehydration as an
energy-efficient process. The anhydrous ethanol production by a pressure
swing adsorption (PSA) process is shown in Figure 8.3. Commonly, a
vapor-phase adsorption (vapor feed) is used to process directly the output
stream from the conventional distillation stage (hydrated ethanol) [16].
Two zeolite adsorption units (A and B) in parallel arrangement are applied
for a continuous ethanol dehydration process. While in one bed (e.g., A) is
operating for dehydrating of superheated ethanol vapor at high pressure,
the other bed (e.g., B) is being regenerated at vacuum conditions by
recirculating (splitter valve C) a small portion of the anhydrous ethanol
through the saturated sieves [17].
Adsorption using molecular sieve for ethanol dehydration can present
a lower energy consumption as compared to azeotropic distillation, since
the stream to be processed only must be vaporized one time. On the other
hand, zeolite is a highly selective material. Water is strongly adsorbed,
requiring high temperatures or a low operating pressure to regenerate the
zeolite bed. In addition, the constant use of superheated vapor for the
regeneration of the sieve bed, accelerates the sieve deterioration due to
thermal shock [18].
Author Copy
a vapor with a higher ethanol composition as compared to a conventional
distillation. In the case of water, its volatility is reduced by the salt effect.
This effect is very effective for the separation of azeotropic mixtures, as in
the case of ethanol-water mixture, promoting a vapor composition of ethanol
higher than the azeotrope.
However, the use of salts as an extraction agent in ED has not reached a
successful application due to the technical problems. The recrystallization of
the salt (recovery step), its dissolution, and the special needs of the construc-
tion materials to avoid corrosion problems have limited the applicability of
dissolved salts as a solvent in ED in the industry [20]. Although the limitation
by technical issues in this technology promotes little interest in the industry,
dissolved salts are a potential alternative to produce anhydrous ethanol at a
low cost.
In Figure 8.4 is shown a simplified scheme of ED with dissolved salts
for ethanol separation from aqueous solutions. The ethanol-water mixture is
fed to the ED column (1), where the separation by distillation is governed
by the modified liquid-vapor equilibrium due to the salt effect. The salt feed
is located immediately after the splitter at the top of the column, where
it is dissolved (2) in the reflux stream [22]. Since the salt is not a volatile
component, it only remains inside the liquid phase and flows downward along
the column. Then, the salt can be recovered at the bottom of the column and
recycled by evaporation or drying, rather than by distillation.
The most common salts used in ED for ethanol dehydration are
CH3COONa, CH3COOK, CaCl2, Ca(NO)2, KI, and NaI [22–24]. In addition,
a mixture of these latter salts can be employed. For instance, a salt with
a ratio 70/30 of CH3COOK/CH3COONa is able to produce ethanol with a
composition of 99.8 wt.%. The energy consumption and capital costs are
lower as compared to AD using benzene as entrainer and conventional
ED with ethylene glycol as solvent [22]. On the other hand, a pilot plant
for producing anhydrous ethanol using CaCl2 was performed [21]. It was
possible to reduce the energy consumption by more than 50% as compared
to conventional ED and vacuum distillation. In other work, different
salts (KCl, CaCl2, NaCl, and KI) were theoretically evaluated to produce
Apple Academic Press
Author Copy
FIGURE 8.4 Typical separation of ethanol-water mixture by extractive distillation using
dissolved salts. Extractive distillation column (1); and salt dissolver (2).
Source: Adapted from: Ref. [21].
Author Copy
In general, the ED with ILs presents a similar configuration of an ED
using conventional solvents. According to Figure 8.5, hydrated ethanol
(azeotropic) and the solvent (ILs) are fed to the ED column (1). The IL is fed
at the top of the column to increase the separation performance. A solvent-
free anhydrous ethanol is produced at the top of the ED column, while the
bottom stream (solvent with water) is partially split by a flash evaporation
(2). The solvent is recovered in the regeneration column (3), where water
is produced at the top of the column, while the recovered solvent from the
bottoms is recycled again to column (1).
FIGURE 8.5 General scheme of an extractive distillation using ionic liquids as solvent.
Extractive distillation column (1); flash drum (2); and solvent regeneration column (3).
Source: Adapted from: Seiler et al. [27].
Several studies have shown the potential of using IL for the ethanol
dehydration. For instance, the liquid-vapor equilibrium for ethanol-water-
solvent (IL) mixture was evaluated close to the azeotropic region [28]. The
ILs used were 1-butyl-3-methylimidazolium bromide ([BMIM][Br]) and
Apple Academic Press
Author Copy
reboilers, and cooler can be achieved using ILs as solvents in comparison to
a conventional solvent ED [30].
Author Copy
8.3.2.1 PERVAPORATION (PV)
PEtOH
θEtOH / w = (1)
Pw
In Eqn. (1), P is the permeability coefficient of the components. P is a
Apple Academic Press
Author Copy
FIGURE 8.6 Sketch of a pervaporation process.
Source: Adapted from: Baker [34].
Author Copy
achieve a similar separation duty of a conventional distillation. It involves a
large membrane area, a meaningful increase in capital cost.
Also, poly(1-trimethylsily-1-propyne) (PTMSP) membranes have shown
potential characteristics to recover ethanol from aqueous solution. In the case
of a PV with PTMSP membranes, the permeate flux and the ethanol selectivity
can be 3 and 2 times higher as compared to PDMS membranes, respectively
[41]. Although PTMSP membranes initially showed a high performance,
these are hardly applied for practical separation due to their instability. For
instance, in a separation of an ethanol aqueous solution (50 wt.%) by PTMSP
membranes, the permeate flux and ethanol selectivity can be reduced up to
60% and 22% after 200 h of operation [42]. PTMSP membranes have been
submitted to different studies in order to improve the stability by modifying
the polymer synthesis, as grafting copolymer to the selective layer. The
synthesis of PTMSP using NbCl5 and TaCl5/Al(i-Bu)3 as catalyst allowed a
stable membrane, even with a low pH PV [43]. In a hybrid PTMSP/PDMS
membranes, a higher permeate flux and ethanol selectivity were achieved as
compared to a solely PTMSP membranes [44]. No membrane instability was
reported. In addition, a permeate with an ethanol composition up to 70 wt.%
was reached from a feed with 7 wt.% of ethanol.
Although PDMS has been the preferred choice to recover ethanol,
inorganic membranes present a substantial improvement in the separation of
ethanol from aqueous solution. Membranes based on hydrophobic zeolites
have shown a higher performance for recovering ethanol in comparison to
PDMS membranes [39]. A special characteristic of zeolite membranes is
a lower swelling as compared to unmodified polymeric membranes, as a
result, a high selectivity can be achieved in PV at high temperatures [45].
MFI membranes (silicate-1 and ZSM-5) are the representative zeolite-based
membranes. It is possible to perform a PV at high temperature with MFI
Author Copy
as compared to azeotropic distillation and other conventional distillation
schemes. A small amount of water is needed to be separated. Consequently,
the operation costs for the dehydration by PV can be significantly lower than
an azeotropic distillation. An installed ethanol dehydration process by PV
was compared to commercial dehydration with AD [48]. The ethanol losses
in PV were neglected, and the alcohol efficiency in PV was about to 99.7%.
On the other hand, the ethanol quality in the AD was lower as compared to
PV. The permeate had not any impurity of the entrainer. Besides, the dehy-
dration by PV presented a reduction of operation costs around 30%. This
reduction was associated with low energy consumption (thermal utilities) in
PV, and the entrainer make up in AD. This latter represented about 10% of
the total operation cost.
For the application of PV and membrane selection in the ethanol dehy-
dration, the liquid-vapor equilibrium and PV performance can be used. In
Figure 8.7 is presented the comparison of the separations between distillation
and PV with different membrane material. A proper dehydration membrane
should be under the 45° line in the liquid-vapor equilibrium of the ethanol-
water mixture. For this case, only the membrane selectivity is considered.
In general, the three membrane types are able to remove water from the
mixture selectively. For hydrated ethanol (≈ 93 wt.%), PVA membranes are
the most appropriate to produce anhydrous ethanol. In the range of 85–95
wt.%, the highest water selectivity is achieved by the PVA membrane. This
reason makes that PVA membrane be widely used in the industry to dehy-
drate ethanol [48].
1. Organic Membranes: As it was mentioned above, PVA-based
membranes are one of the most studied and applied PV membranes
for alcohol dehydration. As a polymeric membrane, PVA is able to be
swelled in ethanol solutions, decreasing its selectivity. However, due
Author Copy
Chitosan (CS) membranes are well-known hydrophilic membranes,
which have been widely applied for ethanol dehydration. However, the
performance of CS membranes can be decreased in aqueous solution
due to its uncontrollable swelling [56]. Similar to PVA, CS membranes
have been researched in order to improve their physical properties.
For instance, a hybrid CS/PVA membrane was synthesized for an
alcohol-water mixture (10 wt.% water) separation at low temperature
[57]. A permeate flux of 0.113 kg m–2 h–1 was achieved and, the water
selectivity was up to 17,000. In other work, a PV with hybrid CS/poly
(sodium vinyl-sulfonate) was performed to dehydrate an ethanol solu-
tion at the azeotropic composition [58]. The experiments were carried
out during 120 h, and the membrane presented a stable performance
with a permeate flux of 1.98 kg m–2 h–1 and a water composition of
99.5 wt.% on the permeate side.
2. Inorganic Membranes: An advantage of inorganic membranes, for
example, ceramic membranes, is the stability and mechanical prop-
erties in PV at high temperatures. Although the membrane selectivity
is naturally given by the membrane material, the permeate flux and
permeate composition are dependent on the vapor pressure of the
component. In the case of a dehydration membrane, the water pres-
sure and, consequently, the water flux can be enhanced by increasing
the feed temperature. Since inorganic membranes do not present
thermal instability, a wider operation range for PV can be presented
as compared to organic membranes.
Microporous silica membrane is one of the most representative inorganic
membranes for alcohol dehydration. A PV with silica membranes for an ethanol-
water mixture (6 wt.% water) can achieve a permeate flux of 0.417 kg m–2 h–1
and a water selectivity of 207 [59]. Moreover, this type of membrane is widely
FIGURE 8.7 Comparison for the separation of ethanol-water mixtures by distillation based on
Author Copy
the liquid-vapor equilibrium (—) and by three pervaporation membranes: (---) cellulose triacetate
(CTA), an anionic polyelectrolyte membrane (– • –) and poly(vinyl alcohol) (PVA) (…).
Source: Adapted from: Ref. [49].
Author Copy
8.3.2.2 HYBRID DISTILLATION-PERVAPORATION (PV) COLUMNS
Author Copy
FIGURE 8.8 Hybrid distillation-pervaporation column to dehydrate ethanol from the
phlegma streams generated in the first column in a conventional distillery.
of the latent heat of this stream. Anhydrous ethanol is produced in the vapor
permeation unit 2 (D). Since the ethanol composition is significantly high
in the feed of D, the permeate generated in this unit (57 wt.% of ethanol) is
condensed and mixed with the wine fed to the separation system. Finally, the
Apple Academic Press
Author Copy
Author Copy
8.3.3 IN SITU ETHANOL RECOVERY: HYBRID FERMENTATION
TECHNOLOGY
Author Copy
selectivity for the solute, and biocompatibility with the yeast. It is desirable
the solvent be low cost, low viscosity, completely immiscible with the
aqueous phase, high selectivity and distribution coefficient for ethanol, and
simple regeneration (for recovery of the product from the solvent phase, in
this case, the ethanol).
Thermodynamically, the distribution coefficient at equilibrium condi-
tion in a liquid-liquid system (equality of chemical potential between both
phases) is defined as the ratio between the equilibrium concentrations of
solute in the organic phase and the aqueous phase [75]. A high distribution
coefficient is convenient because it results in a decrease in the number of
stages of the LLX process and reduces the amount of required solvent for
removal of the desired solute [76, 77]. The polarity of the solvent influences
on the solubility of ethanol. The distribution coefficient of polar solvents
increases as the number of polar groups in the structure of the solvent
increases. The distribution coefficient of ethanol (KD,EtOH) in several solvents
has been widely tested, and in Table 8.1 are shown the highest distribution
coefficient values achieved.
The separation factor, defined as the ratio of the distribution coefficient
of ethanol to water (α = KD,EtOH/KD,water), is used to define the ability of the
solvent to remove the ethanol from water. Values of the separation factor
higher than the unity result in a selective extraction of ethanol from water.
For the selection of the solvent, it is also important to know the selectivity
on other components that are usually within the fermentation broths, specifi-
cally sugars and by-products that may achieve considerable concentrations
during fermentation. Alcohols and esters have shown a high separation factor
for ethanol [78]. Furthermore, the branched-chain solvents have higher
values of the separation factor as compared to its counterpart, the linear-
chain solvents [76]. In many cases, the solvent has shown high selectivity
TABLE 8.1 Solvents with High Values of Distribution Coefficient (around 1 or higher
than 1) for Ethanol Extraction
Author Copy
Solvent KD
1-Hexanol 1.0–1.2
3-Methylcyclo-hexanol 0.93
3-Methyl-3-pentanol 1.3
4-Methyl-2-pebtanol 1.1
2-Ethyl-butanol 0.69–1.03
3-Ethyl-3-pentanol 1.1
Phenol 2.15
o-Isopropylphenol 1.4
o-ter-butylphenol 1.4
Valeric acid 1.3
Hexanoic acid 0.944–1.1
Methyl acetate 0.91
Ethyl propionate 2.53
50% Hexan-1-ol + 50% 2-ethyl-1-ol (w/w) 1.03
Source: Adapted from: Ref. [76].
The stress of the organic solvents on yeast cells affects their morphology
and physiological activity [81, 83]. The accumulation of organic solvents
on the membrane cells modifies their fluidity and permeability [81], which
involves hindrance of nutrients transport, leakage of metabolites, and in the
Apple Academic Press
most severe cases, lysis of the cell [76, 83, 84]. On the other hand, organic
solvent within the cells may disturb the metabolic reactions decreasing the
activity of enzymes because it blocks the enzymatic active sites [81, 83,
84]. As a response to the presence of organic solvents, both the membrane
cell and into the cell, cells may change its lipid and protein composition in
the membrane to modify the plasma membrane properties, such as rigidity,
permeability, and fluidity [84]. However, any response of the cells to the
Author Copy
organic solvent is energy demanding, which means that a portion of the
consumed substrate must be used to supply this energetic demand [79, 83],
decreasing both the cell growth and the metabolite production [76, 79].
Nonpolar solvents are nontoxic to cells [83]. However, nonpolar solvents
provide of low distribution coefficient to ethanol. In general, solvents with a
high distribution coefficient are toxic to cells [76, 79]. For instance, n-alcohols
with high molecular weight have a low distribution coefficient for ethanol,
while n-alcohols with low molecular weight have high toxicity and solubility
in water [70]. According to several works, the efficiency on the in-situ ethanol
removal and toxicity by using yeast cells follows [85, 86]: carboxylic acid >
alcohols > esters > amines > ketones > ethers > hydrocarbons. However, the
branch or length of the linear structure may affect [86].
Natural solvents, for instance, vegetable oils, which are very lipophilic,
are nontoxic to the whole cells [83]. Solvents, such as n-amyl alcohol,
1-octanol, and 1-docecanol have been tested for in situ ethanol removal from
the fermentation broth, resulting n-amyl alcohol the solvent with the highest
recovery percentage, providing the highest ethanol productivity, despite its
toxic effect [72]. Also, n-dodecanol has probed as an efficient solvent for
removal of ethanol by LLX in a continuous fermentation [87, 88].
The toxicity of solvents on the yeast may be reduced and even avoided,
by immobilizing the cells [71, 89]. Several compounds have been tested
for immobilization of cells, such as alginates, collagen, carrageenan, agar,
agarose, glutaraldehyde, and some materials as support such as CS, cellu-
losic material, natural zeolite, and γ-alumina [90, 91]. Calcium-alginate
gel entrapping is the most used method due to it is low cost, provides high
enzymatic activity and simple preparation [90]. Immobilized yeast has
been tested in LLX-fermentation systems for ethanol production, achieving
higher ethanol productivities and yields that its counterpart, the conventional
ethanol fermentation process [72, 76, 88].
Author Copy
FIGURE 8.10 LLX-fermentation arrangements: (a) solvent directly mixed with the fermen-
tation broths inside the fermenter; (b) solvent directly mixed with the fermentation broths in
an external LLX unit; (c) solid porous support contactor inside the fermenter; (d) solid porous
support contactor in an external LLX unit; (e) solvent directly mixed with the fermentation
broths in an external LLX unit with yeast cells filtration; and (f) solid porous support contactor
in an external LLX unit with yeast cells filtration.
Author Copy
The ethanol fermentation also may carry out simultaneously with LLX,
using separate units as Figures 8.10(b, d–f) show. In Figure 8.10(b), the
fermentation broths (B1) from the bioreactor are fed to LLX column. Here,
the solvent phase is in intimate contact with the fermentation broths (ethanol
is transferred from fermentation broth to solvent phase), and both phases are
split in a subsequent unit, the decanter. The aqueous phase from the decanter
(B2) is returned to the bioreactor once is enriched of the substrate with the
fermenter feed (F), while the solvent phase (S1) is fed in the unit for its
regeneration. The clean solvent (S2) and the concentrated ethanol (P) are
produced in the regeneration unit. Finally, the clean solvent (S2) is returned
to the LLX column to perform a new extraction cycle. In this scheme,
immobilization of cells is required in order to avoid the direct contact of the
cells with the solvent phase.
An alternative scheme with a direct yeast-solvent contact is shown in
Figure 8.10(c). However, the mechanism of contacting the fermentation
broths and solvent is significantly different as compared to Figure 8.10(a).
In this case, the yeast cells are directly contacted to the solvent by using
a solid porous support submerged in the fermentation broths. The solid
surface of the support spatially separates the fermentation broths from the
solvent, while the porous, which are filled by the solvent, are exposed to the
fermentation medium. Then, the solid porous support works as contact media
(contactor) between fermentation broths and the solvent phase, providing a
high interfacial area between the phases.
The configuration shown in Figure 8.10(d) is a modification of the
scheme shown in Figure 8.10(c), since the LLX is externally performed
using a contactor (LLX-C). An output stream from the bioreactor is fed
to the LLX-C. In the LLX-C unit, the fermentation broths and solvent are
partially separated by a contactor (solid porous support). Both phases are
contacted through the filled porous of the contactor. Similar to the case of
Figure 8.10(c), the contractor provides a high interfacial area between the
phases, but also, applying an external LLX unit.
The schemes presented in Figure 8.10(e) and f are variants to those previ-
ously showed in Figure 8.10(b) and (d), respectively. The difference lies in
Apple Academic Press
the use of a parallel arrangement of filters before the extraction unit, in order
to avoid direct contact of the yeast cells with the solvent. It is an alternative
for fermentation processes with non-immobilized yeast cells. Due to the
filters suffer from fouling by the yeast cells and solids in the fermentation
broths, two filters must be implemented (F1 and F2). While F1 is operating,
F2 is in a cleaning stage or vice versa.
For all schemes shown in Figure 8.10, the molecular toxicity may be
Author Copy
avoided if there is no recirculation of the fermentation broths (B2). However,
recirculation of the fermentation broth reduces the use of fresh water in the
process significantly, providing a positive impact on the cost of the process
and in the environment. Immobilization of cells works to overcome this
molecular toxicity drawback. Another option is finding a suitable solvent
for ethanol removal, which be nontoxic on the yeast cells. Nowadays, the
seeking of solvents for several processes still is an important field for the
design of new reactive and separation processes. Ionic liquids (ILs), deep
eutectic solvents (DESs) and aqueous two-phase systems (ATPS) are
emerging as potential solvents to be applied in several processes. ILs and
DESs have shown high efficiency for the extraction of several organic acids
[76, 77, 92, 93]. These novel solvents have great potential for the removal
of several metabolites of fermentations. However, it still requires research
on the applicability of these solvents in the ISPR for ethanol fermentations.
Once the solvent is used to remove the ethanol from fermentation broth, it
requires a regeneration step, as was shown in all configurations of the LLX-
fermentation process. There are several choices for the regeneration of the
solvent, such as vacuum flash vaporization, distillation, gas stripping, PV,
and back extraction [76, 77]. Use one or another technique is highly related
to the physicochemical properties of the solvent and ethanol, for example, the
difference in density or differences in volatilities. However, the most used
technique for the regeneration of the solvent rich in ethanol is vacuum flash.
inhibition concentration.
The general scheme of a PV-fermentation is shown in Figure 8.11. In a
hybrid fermentation-PV system, the membrane (C) can be located inside or
outside the fermenter (A). However, an internal membrane is more susceptible
to be fouled due to the high solid concentration (mainly for yeast cells) in
the fermentation broths. As a result, the operating time and membrane perfor-
mance can be rapidly reduced. In order to minimize the membrane fouling due
Author Copy
to the suspended solid, a fraction of the fermentation broths is withdrawn and
passed through a rotatory filter (B). Around 99% of the yeast cell is retained
in B, and recycled to the fermenter again [95]. In the case of a continuous
fermentation, around 85% of retained yeast cells is feedback to the fermenter
again [96]. In the membrane module, ethanol is continuously recovered by a
PV at the fermentation conditions. The retentate is recycled to the fermenter.
Author Copy
[99]. In Figure 8.12 is shown this latter impact of PV on the fermentation
performance. Once PV is initiated, the cell activity arises, increasing the
substrate consumption. Consequently, the ethanol productivity is increased,
allowing an overall ethanol concentration up to 609.8 kg m–3. This is almost
6-fold higher than a conventional fermentation. The reestablishment of the
fermentation performance after the ethanol inhibition condition is reached, is
a general effect of PV on the fermentation process [94, 100, 101].
FIGURE 8.12 Ethanol productivity (—) and glucose consumption rate (…) in a continuous
and closed-circulating fermentation system (CCCF) with PDMS membranes.
Source: Adapted from: Fan et al. [99].
increased due to the increase of its vapor pressure [97]. Furthermore, the
reduction of the membrane selectivity can occur by the presence of the
main fermentation products at a high concentration [97]. Ethanol, acetone,
n-butanol, and 2-propanol at 100 kg m–3, 10 kg m–3, 20 kg m–3 and 2–5 kg
m–3, respectively, reduce the PDMS membrane selectively. Also, the pH
can modify the hydrophobicity of PDMS membranes [103]. At low pH, the
water permeability is increased. The ethanol selectivity is reduced.
Author Copy
KEYWORDS
• chitosan
• extractive distillation
• hyperbranched polymer
• liquid-liquid extraction
• membrane assisted vapor stripping
• polydimethylsiloxane
REFERENCES
1. Zhang, K., Lively, R. P., Noel, J. D., Dose, M. E., McCool, B. A., Chance, R. R., &
Koros, W. J., (2012). Adsorption of water and ethanol in MFI-type zeolites. Langmuir,
28, 8664–8673. doi: 10.1021/la301122h.
2. McAloon, A., Taylor, F., Yee, W., Ibsen, K., & Wooley, R., (2000). Determining the Cost
of Producing Ethanol from Corn Starch and Lignocellulosic Feedstocks. Golden, CO
(United States). doi: 10.2172/766198.
3. Phisalaphong, M., Srirattana, N., & Tanthapanichakoon, W., (2006). Mathematical
modeling to investigate temperature effect on kinetic parameters of ethanol fermentation.
Biochem. Eng. J., 28, 36–43. doi: 10.1016/j.bej.2005.08.039.
4. Lopes, M. L., De Paulillo, S. C. L., Godoy, A., Cherubin, R. A., Lorenzi, M. S., Giometti,
F. H. C., Bernardino, C. D., et al., (2016). Ethanol production in Brazil: A bridge between
science and industry. Brazilian J. Microbiol., 47, 64–76. doi: 10.1016/j.bjm.2016.10.003.
5. Huang, H. J., Ramaswamy, S., Tschirner, U. W., & Ramarao, B. V., (2008). A review of
separation technologies in current and future biorefineries. Sep. Purif. Technol. 62, 1–21.
doi: 10.1016/j.seppur.2007.12.011.
6. Banat, F. A., Al-Rub, F. A. A., & Simandl, J., (2000). Analysis of vapor-liquid equilibrium
of ethanol-water system via headspace gas chromatography: Effect of molecular sieves.
Sep. Purif. Technol., 18, 111–118. doi: 10.1016/S1383-5866(99)00057-X.
7. Dias, M. O. S., Modesto, M., Ensinas, A. V., Nebra, S. A., Filho, R. M., & Rossell, C. E.
Apple Academic Press
Author Copy
10. Widagdo, S., & Seider, W. D., (1996). Journal review: Azeotropic distillation. AIChE J.,
42, 96–130. doi: 10.1002/aic.690420110.
11. Zhao, L., Lyu, X., Wang, W., Shan, J., & Qiu, T., (2017). Comparison of heterogeneous
azeotropic distillation and extractive distillation methods for ternary azeotrope
ethanol/toluene/water separation. Comput. Chem. Eng., 100, 27–37. doi: 10.1016/j.
compchemeng.2017.02.007.
12. Chianese, A., & Zinnamosca, F., (1990). Ethanol dehydration by azeotropic distillation with
a mixed-solvent entrainer. Chem. Eng. J., 43, 59–65. doi: 10.1016/0300-9467(90)80001-S.
13. Bastidas, P., Gil, I., & Rodríguez, G., (2010). Comparison of the main ethanol dehydration
technologies through process simulation. In: 20th Eur. Symp. Comput. Aided Process Eng.
ESCAPE 20.
14. Gomis, V., Pedraza, R., Saquete, M. D., Font, A., & García-Cano, J., (2015). Ethanol
dehydration via azeotropic distillation with gasoline fraction mixtures as entrainers:
A pilot-scale study with industrially produced bioethanol and naphta. Fuel Process.
Technol., 140, 198–204. doi: 10.1016/j.fuproc.2015.09.006.
15. Uyazán, A. M., Gil, I. D., Aguilar, J. L., Rodríguez, G., & Caicedo, L. A., (2004).
Deshidratación del etanol. Ing. e Investig., 49–59.
16. Karimi, S., Tavakkoli, Y. M., & Karri, R. R., (2019). A comprehensive review of
the adsorption mechanisms and factors influencing the adsorption process from the
perspective of bioethanol dehydration. Renew. Sustain. Energy Rev., 107, 535–553. doi:
10.1016/j.rser.2019.03.025.
17. Pedraza, B. R., (2012). Deshidratación de Etanol Mediante Destilación Azeotrópica
con Hidrocarburos Componentes de la Gasolina: Estudio de la Viabilidad del Proceso
a Escala Semi-Planta Piloto. Universidad de Alicante.
18. Jacques, K. A., Lyons, T. P., & Kelsall, D. R., (2003). The Alcohol Textbook: A Reference
for the Beverage, Fuel, and Industrial Alcohol Industries (4th edn.).
19. Gerbaud, V., & Rodriguez-Donis, I., (2014). Extractive distillation. In: Andrzej,
G., & Žarko, O., (eds.), Distillation (1st edn., pp. 201–245). Elsevier. doi: 10.1016/
B978-0-12-386878-7.00006-1.
20. Soares, R. B., Pessoa, F. L. P., & Mendes, M. F., (2015). Dehydration of ethanol with
different salts in a packed distillation column. Process Saf. Environ. Prot., 93, 147–153.
doi: 10.1016/j.psep.2014.02.012.
21. Barba, D., Brandani, V., & Di Giacomo, G., (1985). Hyperazeotropic ethanol salted-out
by extractive distillation. Theoretical evaluation and experimental check. Chem. Eng.
Sci., 40, 2287–2292. doi: 10.1016/0009-2509(85)85130-7.
22. Furter, W. F., (1992). Extractive distillation by salt effect. Chem. Eng. Commun., 116,
35–40. doi: 10.1080/00986449208936042.
23. Ligero, E., & Ravagnani, T. M., (2003). Dehydration of ethanol with salt extractive
distillation—a comparative analysis between processes with salt recovery. Chem. Eng.
Apple Academic Press
Author Copy
27. Seiler, M., Jork, C., Kavarnou, A., Arlt, W., & Hirsch, R., (2004). Separation of azeotropic
mixtures using hyperbranched polymers or ionic liquids. AIChE J., 50, 2439–2454. doi:
10.1002/aic.10249.
28. Tsanas, C., Tzani, A., Papadopoulos, A., Detsi, A., & Voutsas, E., (2014). Ionic liquids
as entrainers for the separation of the ethanol/water system. Fluid Phase Equilib., 379,
148–156. doi: 10.1016/j.fluid.2014.07.022.
29. Zhu, Z., Ri, Y., Li, M., Jia, H., Wang, Y., & Wang, Y., (2016). Extractive distillation
for ethanol dehydration using imidazolium-based ionic liquids as solvents. Chem. Eng.
Process. - Process Intensif., 109, 190–198. doi: 10.1016/j.cep.2016.09.009.
30. Aniya, V., De, D., Singh, A., & Satyavathi, B., (2018). Design and operation of
extractive distillation systems using different class of entrainers for the production of
fuel-grade tert-butyl alcohol: A techno-economic assessment. Energy, 144, 1013–1025.
doi: 10.1016/j.energy.2017.12.099.
31. Seiler, M., Köhler, D., & Arlt, W., (2002). Hyperbranched polymers: New selective
solvents for extractive distillation and solvent extraction. Sep. Purif. Technol., 29,
245–263. doi: 10.1016/S1383-5866(02)00163-6.
32. Seiler, M., Buggert, M., Kavarnou, A., & Arlt, W., (2003). From alcohols to
hyperbranched polymers: The influence of differently branched additives on the vapor−
liquid equilibria of selected azeotropic systems. J. Chem. Eng. Data, 48, 933–937. doi:
10.1021/je025644w.
33. Seiler, M., Arlt, W., Kautz, H., & Frey, H., (2002). Experimental data and theoretical
considerations on vapor-liquid and liquid-liquid equilibria of hyperbranched
polyglycerol and PVA solutions. Fluid Phase Equilib., 201, 359–379. doi: 10.1016/
S0378-3812(02)00082-1.
34. Baker, R. W., (2012). Membrane Technology and Applications (3rd edn.). Wiley & Sons,
West Sussex.
35. León, J. A., & Fontalvo, J., (2019). PDMS modified membranes by 1-dodecanol and
its effect on ethanol removal by pervaporation. Sep. Purif. Technol., 210, 364–370. doi:
10.1016/j.seppur.2018.08.019.
36. Wijmans, J. G. H., & Baker, R. W., (2006). The solution-diffusion model: A unified
approach to membrane permeation. In: Mater. Sci. Membr. Gas Vap. Sep. (pp. 159–189).
John Wiley & Sons, Ltd, Chichester, UK. doi: 10.1002/047002903X.ch5.
37. Baker, R. W., Wijmans, J. G., & Huang, Y., (2010). Permeability, permeance and
selectivity: A preferred way of reporting pervaporation performance data. J. Membr.
Sci., 348, 346–352.
38. Andre, A., Nagy, T., Toth, A. J., Haaz, E., Fozer, D., Tarjani, J. A., & Mizsey, P., (2018).
Distillation contra pervaporation: Comprehensive investigation of isobutanol-water
separation. J. Clean. Prod., 187, 804–818. doi: 10.1016/j.jclepro.2018.02.157.
39. Peng, P., Shi, B., & Lan, Y., (2010). Review of membrane materials for ethanol recovery
Apple Academic Press
Author Copy
Pervaporation of 50 wt % ethanol-water mixtures with poly(1-trimethylsilyl-1-propyne)
membranes at high temperatures. J. Appl. Polym. Sci., 103, 2843–2848. doi: 10.1002/
app.25375.
43. Volkov, V. V., Fadeev, A. G., Khotimsky, V. S., Litvinova, E. G., Selinskaya, Y.
A., McMillan, J. D., & Kelley, S. S., (2004). Effects of synthesis conditions on the
pervaporation properties of poly[1-(trimethylsilyl)-1-propyne] useful for membrane
bioreactors. J. Appl. Polym. Sci., 91, 2271–2277. doi: 10.1002/app.13358.
44. Nagase, Y., Ishihara, K., & Matsui, K., (1990). Chemical modification of poly(substituted-
acetylene): II. Pervaporation of ethanol / water mixture through poly(1-trimethylsilyl-
1-propyne) / poly(dimethylsiloxane) graft copolymer membrane. J. Polym. Sci. Part B
Polym. Phys., 28, 377–386. doi: 10.1002/polb.1990.090280309.
45. Korelskiy, D., Leppäjärvi, T., Zhou, H., Grahn, M., Tanskanen, J., & Hedlund, J.,
(2013). High flux MFI membranes for pervaporation. J. Memb. Sci., 427, 381–389. doi:
10.1016/j.memsci.2012.10.016.
46. Soydaş, B., Dede, Ö., Çulfaz, A., & Kalıpçılar, H., (2010). Separation of gas and
organic/water mixtures by MFI type zeolite membranes synthesized in a flow system.
Microporous Mesoporous Mater., 127, 96–103. doi: 10.1016/j.micromeso.2009.07.004.
47. Caro, J., Noack, M., & Kölsch, P., (2005). Zeolite membranes: From the laboratory scale
to technical applications. Adsorption, 11, 215–227. doi: 10.1007/s10450-005-5394-9.
48. Sander, U., & Soukup, P., (1988). Design and operation of a pervaporation plant for
ethanol dehydration. J. Memb. Sci., 36, 463–475. doi: 10.1016/0376-7388(88)80036-X.
49. Kaschemekat, J., Barbknecht, B., & Böddeker, K. W., (1986). Konzentrierung von ethanol
Durch pervaporation. Chemie Ing. Tech., 58, 740–742. doi: 10.1002/cite.330580909.
50. Samei, M., Mohammadi, T., & Asadi, A. A., (2013). Tubular composite PVA ceramic
supported membrane for bio-ethanol production. Chem. Eng. Res. Des., 91, 2703–2712.
doi: 10.1016/j.cherd.2013.03.008.
51. Namboodiri, V. V., & Vane, L. M., (2007). High permeability membranes for the
dehydration of low water content ethanol by pervaporation. J. Memb. Sci., 306, 209–215.
doi: 10.1016/j.memsci.2007.08.050.
52. Penkova, A. V., Dmitrenko, M. E., Savon, N. A., Missyul, A. B., Mazur, A. S.,
Kuzminova, A. I., Zolotarev, A. A., et al., (2018). Novel mixed-matrix membranes based
on polyvinyl alcohol modified by carboxy fullerene for pervaporation dehydration. Sep.
Purif. Technol., 204, 1–12. doi: 10.1016/j.seppur.2018.04.052.
53. Castro-Muñoz, R., Buera-González, J., De La Iglesia, O., Galiano, F., Fíla, V.,
Malankowska, M., Rubio, C., et al., (2019). Towards the dehydration of ethanol using
pervaporation cross-linked poly(vinyl alcohol)/graphene oxide membranes. J. Memb.
Sci., 582, 423–434. doi: 10.1016/j.memsci.2019.03.076.
Apple Academic Press
54. Dmitrenko, M. E., Penkova, A. V., Kuzminova, A. I., Morshed, M., Larionov, M. I.,
Alem, H., Zolotarev, A. A., et al., (2018). Investigation of new modification strategies
for PVA membranes to improve their dehydration properties by pervaporation. Appl.
Surf. Sci., 450, 527–537. doi: 10.1016/j.apsusc.2018.04.169.
55. Yuan, H., Wang, C., Liu, X., & Lu, J., (2020). Preparation of PVA-PFSA-Si pervaporative
hybrid membrane and its dehydration performance. Polym. Bull. doi: 10.1007/
s00289-020-03107-5.
56. Jyothi, M. S., Reddy, K. R., Soontarapa, K., Naveen, S., Raghu, A. V., Kulkarni, R. V.,
Suhas, D. P., et al., (2019). Membranes for dehydration of alcohols via pervaporation. J.
Author Copy
Environ. Manage, 242, 415–429. doi: 10.1016/j.jenvman.2019.04.043.
57. Rao, K. S. V. K., Subha, M. C. S., Sairam, M., Mallikarjuna, N. N., & Aminabhavi,
T. M., (2007). Blend membranes of chitosan and poly(vinyl alcohol) in pervaporation
dehydration of isopropanol and tetrahydrofuran. J. Appl. Polym. Sci., 103, 1918–1926.
doi: 10.1002/app.25078.
58. Zheng, P. Y., Ye, C. C., Wang, X. S., Chen, K. F., An, Q. F., Lee, K. R., & Gao, C. J.,
(2016). Poly(sodium vinyl sulfonate)/chitosan membranes with sulfonate ionic cross-
linking and free sulfate groups: Preparation and application in alcohol dehydration. J.
Memb. Sci., 510, 220–228. doi: 10.1016/j.memsci.2016.02.060.
59. Duque, S. A. C., Gómez, G. M. Á., Fontalvo, J., Jedrzejczyk, M., Rynkowski, J. M., &
Dobrosz-Gómez, I., (2013). Ethanol dehydration by pervaporation using microporous silica
membranes. Desalin. Water Treat, 51, 2368–2376. doi: 10.1080/19443994.2012.728053.
60. Peters, T. A., Fontalvo, J., Vorstman, M. A., Benes, N. E., Van, D. R. A., Vroom, Z. A. E.
P., Soest-Vercammen, L. J. V., & Keurentjes, J. T. F., (2005). Hollow fibre microporous
silica membranes for gas separation and pervaporation synthesis, performance and
stability. J. Membr. Sci., 248, 73–80.
61. Sairam, M., Patil, M., Veerapur, R., Patil, S., & Aminabhavi, T., (2006). Novel dense
poly(vinyl alcohol)–TiO2 mixed matrix membranes for pervaporation separation of
water-isopropanol mixtures at 30°C. J. Memb. Sci., 281, 95–102. doi: 10.1016/j.
memsci.2006.03.022.
62. Yang, D., Li, J., Jiang, Z., Lu, L., & Chen, X., (2009). Chitosan/TiO2 nanocomposite
pervaporation membranes for ethanol dehydration. Chem. Eng. Sci., 64, 3130–3137.
doi: 10.1016/j.ces.2009.03.042.
63. Sudhakar, H., Venkata, P. C., Sunitha, K., Chowdoji, R. K., Subha, M. C. S., & Sridhar, S.,
(2011). Pervaporation separation of IPA-water mixtures through 4A zeolite-filled sodium
alginate membranes. J. Appl. Polym. Sci., 121, 2717–2725. doi: 10.1002/app.33695.
64. Ma, N., Wang, R., He, G., & Wang, Z., (2018). Preparation of high-performance zeolite
NaA membranes in clear solution by adding SiO2 into Al2O3 hollow-fiber precursor.
AIChE J., 64, 2679–2688. doi: 10.1002/aic.16107.
65. Huang, Y., Baker, R. W., & Vane, L. M., (2010). Low-energy distillation-membrane
separation process. Ind. Eng. Chem. Res., 49, 3760–3768. doi: 10.1021/ie901545r.
66. Kunnakorn, D., Rirksomboon, T., Siemanond, K., Aungkavattana, P., Kuanchertchoo,
N., Chuntanalerg, P., Hemra, K., et al., (2013). Techno-economic comparison of energy
usage between azeotropic distillation and hybrid system for water-ethanol separation.
Renew. Energy, 51, 310–316. doi: 10.1016/j.renene.2012.09.055.
67. Roza, M., & Maus, E., (2006). Industrial experience with hybrid distillation-pervaporation
or vapor permeation applications. Inst. Chem. Eng. Symp. Ser., 152, 619–627.
Apple Academic Press
68. Vane, L. M., & Alvarez, F. R., (2015). Effect of membrane and process characteristics
on cost and energy usage for separating alcohol-water mixtures using a hybrid vapor
stripping-vapor permeation process. J. Chem. Technol. Biotechnol., 90, 1380–1390. doi:
10.1002/jctb.4695.
69. Woodley, J. M., Bisschops, M., Straathof, A. J. J., & Ottens, M., (2008). Future directions
forin-situ product removal (ISPR). J. Chem. Technol. Biotechnol., 83, 121–123. doi:
10.1002/jctb.1790.
70. Díaz, M., (1988). Three-phase extractive fermentation. Trends Biotechnol., 6, 126–130.
doi: 10.1016/0167-7799(88)90102-3.
Author Copy
71. Zentou, H., Abidin, Z. Z., Yunus, R., Biak, D. R. A., & Korelskiy, D., (2019). Overview
of alternative ethanol removal techniques for enhancing bioethanol recovery from
fermentation broth. Processes, 7, 1–16. doi: 10.3390/pr7070458.
72. Widjaja, T., Altway, A., Permanasari, A. R., & Gunawan, S., (2014). Production of ethanol
as a renewable energy by extractive fermentation. Appl. Mech. Mater., 493, 300–305. doi:
10.4028/www.scientific.net/AMM.493.300.
73. Honda, H., Taya, M., & Kobayashi, T., (1986). Ethanol fermentation associated with
solvent extraction using immobilized growing cells of saccharomyces cerevisiae and its
lactose-fermentable fusant. J. Chem. Eng. Japan, 19, 268–273. doi: 10.1252/jcej.19.268.
74. Kollerup, F., & Daugulis, A. J., (1986). Ethanol production by extractive fermentation-
solvent identification and prototype development. Can. J. Chem. Eng., 64, 598–606.
doi: 10.1002/cjce.5450640410.
75. Sandler, S. I., (2006). Chemical, Biochemical, and Engineering Thermodynamics (4th
edn.). John Wiley & Sons Inc.
76. Kim, J. K., Iannotti, E. L., & Bajpai, R., (1999). Extractive recovery of products from
fermentation broths. Biotechnol. Bioprocess Eng., 4, 1–11. doi: 10.1007/BF02931905.
77. Vane, L. M., (2008). Separation technologies for the recovery and dehydration of alcohols
from fermentation broths. Biofuels, Bioprod. Biorefining, 2, 553–588. doi: 10.1002/
bbb.108.
78. Matsumura, M., & Märkl, H., (1986). Elimination of ethanol inhibition by perstraction.
Biotechnol. Bioeng., 28, 534–541. doi: 10.1002/bit.260280409.
79. Pérez, A. D., Rodríguez-Barona, S., & Fontalvo, J., (2018). Molecular toxicity of
potential liquid membranes for lactic acid removal from fermentation broths using
Lactobacillus casei ATCC 393. Dyna, 85, 360–366. doi: 10.15446/dyna.v85n207.72374.
80. Pérez, A. D., Gómez, V. M., Rodríguez-Barona, S., & Fontalvo, J., (2019). Liquid-liquid
equilibrium and molecular toxicity of active and inert diluents of the organic mixture
tri-iso-octylamine/dodecanol/dodecane as potential membrane phase for lactic acid
removal. J. Chem. Eng. Data.
81. Xu, K., Lee, Y. S., Li, J., & Li, C., (2019). Resistance mechanisms and reprogramming
of microorganisms for efficient biorefinery under multiple environmental stresses.
Synth. Syst. Biotechnol., 4, 92–98. doi: 10.1016/j.synbio.2019.02.003.
82. Marták, J., Sabolová, E., Schlosser, Š., Rosenberg, M., & Kristofíková, L., (1997).
Toxicity of organic solvents used in situ in fermentation of lactic acid by Rhizopus
arrhizus. Biotechnol. Tech., 11, 71–75. doi: 10.1023/A:1018408220465.
83. Heipieper, H. J., Weber, F. J., Sikkema, J., Keweloh, H., & De Bont, J. A. M., (1994).
Mechanisms of resistance of whole cells to toxic organic solvents. Trends Biotechnol.,
12, 409–415. doi: 10.1016/0167-7799(94)90029-9.
84. Dyrda, G., Boniewska-Bernacka, E., Man, D., Barchiewicz, K., & Słota, R., (2019). The
Apple Academic Press
Author Copy
solvent extraction. Biotechnol. Lett., 3, 405–408. doi: 10.1007/BF01134098.
88. Gyamerah, M., & Glover, J., (1996). Production of ethanol by continuous fermentation
and liquid-liquid extraction. J. Chem. Technol. Biotechnol., 66, 145–152. doi: 10.1002/
(SICI)1097-4660(199606)66:2<145::AID-JCTB484>3.0.CO;2-2.
89. Kapucu, H., & Mehmetoǧlu, Ü., (1998). Strategies for reducing solvent toxicity in
extractive ethanol fermentation. Appl. Biochem. Biotechnol. - Part A Enzym. Eng.
Biotechnol., 75, 205–214. doi: 10.1007/BF02787775.
90. Santos, E. L. I., Rostro-Alanís, M., Parra-Saldívar, R., & Alvarez, A. J., (2018). A novel
method for bioethanol production using immobilized yeast cells in calcium-alginate
films and hybrid composite pervaporation membrane. Bioresour. Technol., 247,
165–173. doi: 10.1016/j.biortech.2017.09.091.
91. Ivanova, V., Petrova, P., & Hristov, J., (2011). Application in the ethanol fermentation
of immobilized yeast cells in matrix of alginate/magnetic nanoparticles, on chitosan-
magnetite microparticles and cellulose-coated magnetic nanoparticles. Int. Rev. Chem.
Eng., 3, 289–299.
92. Vanda, H., Dai, Y., Wilson, E. G., Verpoorte, R., & Choi, Y. H., (2018). Green solvents
from ionic liquids and deep eutectic solvents to natural deep eutectic solvents. Comptes.
Rendus. Chim., 21, 628–638. doi: 10.1016/j.crci.2018.04.002.
93. López-Porfiri, P., Gorgojo, P., & Gonzalez-Miquel, M., (2020). Green solvent selection
guide for biobased organic acid recovery. ACS Sustain. Chem. Eng., 8, 8958–8969. doi:
10.1021/acssuschemeng.0c01456.
94. Leon, J., Palacios-Bereche, R., & Nebra, S. A., (2016). Batch pervaporative fermentation
with coupled membrane and its influence on energy consumption in permeate recovery
and distillation stage. Energy, 109, 77–91.
95. Bassetto, N. Z., (2006). SEPPA-Specialist System for Ethanol Production Plant. Univer-
sidade Estadual de Campinas.
96. Leon, J. A., (2015). Análise do desempenho de membranas de pervaporação no
processo convencional de fermentação para produção de etanol. Universidade Federal
do ABC.
97. Vane, L. M., (2005). A review of pervaporation for product recovery from biomass
fermentation processes. J. Chem. Technol. Biotechnol., 80, 603–629. doi: 10.1002/
jctb.1265.
98. Cho, C. W., & Hwang, S. T., (1991). Continuous membrane fermentor separator for
ethanol fermentation. J. Memb. Sci., 57, 21–42. doi: 10.1016/S0376-7388(00)81160-6.
99. Fan, S., Xiao, Z., Zhang, Y., Tang, X., Chen, C., Li, W., Deng, Q., & Tao, P., (2014).
Enhanced ethanol fermentation in a pervaporation membrane bioreactor with the
convenient permeate vapor recovery. Bioresour. Technol., 155, 229–234.
100. Kargupta, K., Datta, S., & Sanyal, S. K., (1998). Analysis of the performance of a
Apple Academic Press
Author Copy
pH, CO2, and high glucose concentrations on polydimethylsiloxane pervaporation
membranes for ethanol removal. Ind. Eng. Chem. Res., 51, 9328–9334.
Author Copy
MARÍA ALEJANDRA SÁNCHEZ-MUÑOZ,4
HÉCTOR HUGO MOLINA CORREA,4
CRISTIAN EMANUEL GÁMEZ-ALVARADO,4
PERLA ARACELI MELÉNDEZ-HERNÁNDEZ,3 and
JAVIER ULISES HERNÁNDEZ-BELTRÁN4
1
Bioprocesses and Bioprospecting Laboratory, Biological Sciences Faculty,
Autonomous University of Coahuila, Torreón–27000, Coahuila, México
Faculty of Mechanical and Electrical Engineering, Autonomous
2
ABSTRACT
using yeast or bacteria in the final fermentation step. In these three steps,
there are many challenges to face as the conversion of a high concentration
of raw materials, saving chemical and biological reagents, and the reduction
of the processing time. Therefore, it brings into play several engineering
Apple Academic Press
Author Copy
9.1 INTRODUCTION
The fast growth of the human population has increased the demand for
energy and consequently bringing into play concerns as global warming
(GW) and reduction of natural sources [1]. One of the main aspects that the
world faces is the high consumptions of fossil fuels which represent around
80% of the global energy usage [2]. Thus, the R&D in renewable sources of
energy has become a priority. Biofuels are produced from biomass and have
been evaluated as an alternative to reduce both the dependence of fossil fuels
and environmental pollution [3]. In recent decades, they are emerged as an
effective strategy to deal with the matter related to the decreasing of fossil
fuels and the increasing of greenhouse gas (GHG) emissions [4]. Biofuels
can be classified in terms of generation according to the raw material used.
First generation biofuels are derived from food crops which unfortunately
creates ethical concerns and the increase of food prices [5]. To solve these
issues, the second-generation biofuels are produced from inedible LCB
such as agro-residues, forestry wastes, herbaceous biomass, wood wastes or
municipal wastes [6]. Among the most promising biofuels, the bioethanol of
second generation is considered as a fuel substitute for road transportation
due to it is an eco-friendly product with a higher octane number and emits
fewer concentrations of GHGs than gasoline [7].
The LCB is converted into bioethanol through the steps of pretreatment,
enzymatic hydrolysis, and fermentation that usually are carried out in a
bioreactor which is a controlled reaction system where an optimum environ-
ment is provided to enable the highest possible yield [8]. Inherently, different
configurations and operation modes of bioreactors have been explored,
by studying the impact of different process conditions involved such as
substrate concentration, catalyst concentration, temperature, pH, reaction
volume, mixing, among others [9]. A bioreactor commonly used is the stirred
tank reactor (STR) that can be operated under the different modes batch,
fed-batch or continuous, and other innovative bioreactor designs have been
proposed to process lignocellulosic materials [10]. It draws the convenience
to consider the design and engineering parameters involved in the bioreactors
Apple Academic Press
Author Copy
Lignocellulosic biomass is an attractive renewable feedstock for bioethanol
production due to its huge abundance, low cost and sustainable supply [11].
Lignocellulose is composed by three main biopolymers such as cellulose,
hemicellulose, and lignin [12]. Table 9.1 shows the chemical composition for
different LCB. In general, the fractions of cellulose, hemicellulose, and lignin
corresponds to 25–55%, hemicellulose 5–35% and 5–30%, respectively.
Author Copy
the ramifications of the main chains make the hemicellulose structure easier
to hydrolyze than cellulose [19].
Lignin is the most abundant phenolic biopolymer in nature and is the
third major component in lignocellulosic materials. Lignin is related to both
hemicellulose and cellulose, forming a physical seal that is an impenetrable
barrier in the plant cell wall. It is present in the cell wall to provide
structural support, impermeability, and resistance against microbial attack
and oxidative stress. Lignin is an optically inactive, non-water soluble,
amorphous heteropolymer that is formed from phenylpropane bound
through non-hydrolysable bonds. This polymer is generally synthesized by
three different phenylpropanoid derivatives: coniferyl (guaiacyl propanol),
coumaric (p-hydroxyphenyl propanol) and sinapyl (syringyl propanol)
alcohols; synthesized from phenylalanine through various derivatives of
cinnamic acid. Phenylpropanoid alcohols are bound in a polymer by the
action of enzymes that generate intermediates in the form of free radicals.
This heterogeneous structure is linked by C-C and aryl-ether bonds linked to
β-aryl-aryl-glycerol ether, these being the predominant structures [20–22].
Figure 9.1 shows the general process for bioethanol production from LCB, in
which three main steps appear: (1) pretreatment methods have been identified
as the first barrier in is used to break the lignin wall and disrupt the crystallinity
of cellulose to allow the accessibility of the enzymes; (2) enzymatic
hydrolysis process consists in the conversion of holocellulose (cellulose plus
hemicellulose) to reducing sugars through the action of enzymes on pretreated
biomass; (3) fermentation is the step in which the reducing sugars like glucose
is converted to ethanol by using yeast or bacteria [23].
Author Copy
FIGURE 9.1 Outline of bioethanol production process.
9.1.2.1 PRETREATMENT
Author Copy
cost but also maximizes the use of cellulose and hemicellulose [28]. Alkaline-
oxidative is an environment-friendly pretreatment that can be carried out
at mild temperature and pressure, which leads to the formation of minor
inhibitors [16]. Regardless of the type of LCB, if ethanol is the final product,
the pretreatment must guarantee a maximum recovery of holocellulose to
achieve high concentration of reducing sugars. Furthermore, pretreatment
has been identified as the crucial step, both technically and economically, in
the conversion of global LCB due to its strong influences to restrict the other
steps in the chain of bioethanol production.
Author Copy
mixing conditions, and solid-liquid ratio that will be further discussed.
however, this aspect affects directly to the mixing of the reaction and simul-
taneously to the process conditions such as pH and temperature cannot be
constant along the time, even the adjustment of these parameters is difficult
[34]. Thus, there is no possible to ensure the best performance considering
that the mixing trouble begin from 7% w v–1 of substrate concentration [14].
Therefore, the research activities in enzymatic hydrolysis have been focused
to reach high solids concentration of substrate between 20–40% w v–1 using
Author Copy
engineering strategies as fed-batch or semi-continuous configurations for the
reaction instead of a batch one and to deeply study the relationship of the
several parameters which affect the enzymatic hydrolysis process to increase
the profitability of fermentable sugars production [35].
9.1.2.3 FERMENTATION
The reactor configuration, process conditions, and the operation mode during
the pretreatment conversion process are essential to gap the high yields and
productivity in the biomass conversion at industrial levels [40]. More recently,
different studies have been conducted in several bioreactor configurations
with the aim to obtain higher yields of glucose in the pretreatment stage [8].
Author Copy
The pretreatment step has many methods for its development and each
one present their own phenomenology into the reaction. The mechanical
method is the most used among the physical pretreatments [41] but does
not require a bioreactor for reduction of particle size but rather a milling
equipment, however, the reduction of particle size is a process condition that
can limit the performance of the whole bioethanol production chain [42].
For this reason, a study of the optimum particle size for each LCB is a key
issue [26]. Other common physical pretreatments are the microwave and
hydrothermal methods. Microwave reactors (MWR) have noticed interest
for lignocellulosic pretreatment. Microwaves are a kind of electromagnetic
radiation that is shaped like energy propagating in a vacuum in the absence
of any material in motion. Microwaves are originated by the separation of
two charges positive and negative of equal magnitude which are separated by
a fixed distance. MWR are mainly composed by a generator, a cavity section,
a waveguide, and an advanced control monitoring system. The principle
of these reactors is based on Maxwell’s equation [43]. Hermiati et al. [44]
studied microwave-assisted acid pretreatment of sugarcane trash, with a
maximum work temperature of 180°C and biomass capacity of 100 mL. The
sugar production from the pretreated LCB was 20.2 g L–1 of xylose, which
offers a good yield scale under these conditions. In addition, Peng et al. [45]
designed a pilot-scale continuous microwave irradiation reactor (MWIR).
The temperature range of operation is between 0 to 300C, with a maximum
power of 6 kW. MWIR ethanol yield indicated a value of 31.2 g 100 g–1 of
pretreated corn straw. Table 9.2 describes the main characteristics of reactors
system for conversion of LCB considering the reaction system, including the
agitation method, reaction volume, reactor configuration, and the reducing
sugars or ethanol yield. On the other hand, an emerging technology to convert
wet biomass into valuable products due to the action of temperature (200–
370°C) and high pressure (4–20 MPa) is hydrothermal liquefaction reactors
TABLE 9.2 Main Characteristics of Reactors System for Conversion of Lignocellulosic Biomass
Biomass Reactor System Agitation Tank Configuration Sugarsa or Ethanolb References
Volume Process Concentration Yield
b
Agave bagasse Tubular reactor Rushton-type stirred 1.0 L Batch SSF 37.6 g L–1 [116]
blades
Sugarcane bagasse Hydrodynamic cavitation Recirculation pump 2.0 L Fed-batch SSF b28.44 g L–1 [48]
b –1
Dry corn Hydrodynamic cavitation Low-shear impeller 379 L Batch SSF 35.8% w w [49]
b
Reed grass Hydrodynamic cavitation Recirculation pump 150 ml Batch SSF 25.9 g L–1 [50]
a –1
Apple pomace Bubble column bioreactor Air bubbles – Batch 51.8 g L fermentable sugar [117]
a
Agave bagasse High pressure stirred reactor Anchor impeller 1.0 L Batch 3.7 g L–1 xylose [8]
a
Sugarcane bagasse Plasma in liquid reactor N/A 1.0 L Batch 51.3% glucose [118]
b –1
Sargassum horneri Stirred tank bioreactor Rushton impeller 2.0 L Batch S-SSF 2.89 g L [119]
a
Sugarcane bagasse Supercritical CO2 reactor CO2 bubbles 100 ml Batch 1.17 g L–1 fermentable sugars [120]
b
Tobacco stalk Rotating reactor Rotation 2L Batch 27.5 g Kg–1 TS [121]
b
Sugarcane bagasse Fluidized bed reactor Air bubbles 2L Batch SSCF 0.34 g g–1 [122]
b –1
Birch biomass Steam explosión react N/A 4L Batch SSF 80 g L [123]
a
Wheat Straw High pressure reactor Rotation 1L Batch 72.7% glucose [21]
a –1
Sugarcane trash Microwave reactor N/A 100 ml Batch 20.2 g L xylose [44]
Author Copy
Design and Engineering Parameters of Bioreactors 243
(HTLR) [46]. Several studies have revealed the potential of HTLR for bagasse
pretreatment to produce a liquid product, known as bio-oil. Miliotti et al. [47]
evaluated the valorization of lignin-rich stream from lignocellulosic ethanol
production at industrial scale using hydrothermal liquefaction. The stainless-
Apple Academic Press
steel reactor was operated in batch mode with a volume capacity of 27 mL.
It has been observed that reactors configuration and operation mode, in order
to improve the LCB pretreatments, should consider the development of new
technologies and novel methods to overcome the lab-scale into industrial
level. Special attention might be focused on the energy reduction process and
the improvement of sugars-bioethanol yields.
The design of bioreactors in chemical and physicochemical pretreat-
Author Copy
ments differs in the temperature parameter. If the temperature is moderate,
between 25–70°C, stirred tank bioreactors (STBR) are normally used. If the
temperature to be used is higher, for example, more than 100°C, the bioreac-
tors used also tend to withstand high pressures. The substrate concentration
plays an important role to determine the type of stirring and can vary from a
moderate concentration of 5% to a high concentration of >12% (w v–1) [51].
This entails to the use of mechanical mixing and the increasing of speed
of the stirring. The more concentration, the faster the stirring speed, which
leads to higher energy requirement. For this reason, the engineering strate-
gies used for processing high concentrations of substrate in chemical and
physicochemical pretreatments are fed-batch or semi-batch and continuous
modes of operation [10]. Besides, bioreactors must withstand to acid and
alkaline values of, i.e., the value of pH using an acid reagent like sulfuric
acid is <1 or by using an alkaline reagent like sodium hydroxide is >11 [27].
Another physical pretreatment but also can be performed as physio-
chemical one is hydro-dynamic cavitation. It involves the generation of
cavitation in freely flowing liquid which is constricted with a venture tube
or orifice plate [52]. Cavitation phenomenon can be defined as the formation
and subsequent growing and collapse of microbubbles due to the changes in
pressure at constant temperature. Bernoulli's equation is the main principle in
this technique, where the drop in pressure at the constriction below the vapor
pressure of the flowing liquid generates bubbles or hydrodynamic cavities
[53]. Hydrodynamic cavitation reactors (HCR) have been effectively applied
for delignification of several LCB, such as reed grass [50], sugarcane bagasse
[48], dry mill corn [49] among others. The great advantage of HCR is its much
lower energy consumption compared with ultrasonic reactor, as well it has
been observed that this pretreatment affects both delignification and crystal-
linity of cellulose to increase yields in bioethanol production. Terán et al. [48]
Also, in the system was included two tanks for lignocellulosic biomass
preparation and for the cavitation devices consisted in a plate of 8 to 24
orifices at different configuration. On the other hand, Ramirez-Cadavid et
al. [49] built a pilot scale HCR. The system included a 379 L heat-jacketed
vessel able to maintain temperatures from 82°C to 94°C. The agitation was
provided by a low shear impeller mixer.
Author Copy
9.3 BIOREACTORS FOR ENZYMATIC HYDROLYSIS PROCESS
the liquid volume of the reactor keeps constant. However, this process can be
affected by substrate inhibition [60].
Different types of configurations in the bioreactor fields had been employed
including well-known configurations as STBR, which is the most commonly
Apple Academic Press
used for enzymatic processes, and the last advances on STRB which focus
on the conversion of energy crops and residual materials into sugars and
ethanol include separate hydrolysis and fermentation (SHF), simultaneous
saccharification and fermentation (SSF), simultaneous saccharification and
co-fermentation (SSCF) and consolidated bioprocessing (CBP) (Figure 9.3).
Author Copy
FIGURE 9.3 Different types of bioreactor configurations for sugars and bioethanol production.
[Abbreviations: SHF: separate hydrolysis and fermentation; SSF: simultaneous saccharification
and fermentation; SSCF: simultaneous saccharification and co-fermentation; CPB: consolidated
bioprocessing].
Author Copy
enzymes than other simultaneous processes [62, 63].
SSF combines enzymatic hydrolysis and fermentation to obtain value
added products in a single step, which can reduce significatively the time
and cost of the process. Another advantage is the risk reduction of inhibitory
compounds of saccharification. Hence, SSF has been widely used for produc-
tion of biofuels from lignocellulosic materials. However, in comparison with
SHF pH and temperature parameters cannot be individually set and an equi-
librium point might be necessary for the process to work properly [64, 65].
SSCF is referred to the process where enzymatic hydrolysis can be
conducted simultaneously with the co-fermentation of glucose and xylose
in a single process. This process take advantage of the biochemical affinities
due glucose and xylose compete for the same transport systems in the
co-fermentation of xylose. Thus, SSCF is promising for ethanol production
[66]. SSCF have some advantages over SHF which includes the removal of
hydrolysis end-products, short processing time and low contamination risk.
Besides, it has been proved its high efficiency and rates of ethanol yield [67].
CBP is a system where the enzyme production, substrate hydrolysis and
fermentation are performed in one step by lignocellulolytic microorganisms.
The production costs by using this process are lower for biofuel obtention,
due implies lower energy inputs and higher conversion efficiencies in
comparison to SHF. However, CBP has certain disadvantages, mainly
involved with enzyme production because monocultures have shown some
limitations. While co-cultures might have different enzyme production
requirements due bacteria can ferment LCB at mesophilic or thermophilic
temperatures and thus, different rates of cellulose hydrolysis and biofuel
production can be observed [68, 69].
Jacket or coil for heating/cooling system is one of the most important
instruments for laboratory-scale bioreactors, the vessel can be positioned
into a thermal incubator or can be adapted to an electrical heater resistance,
also the external wall of the vessel can be equipped with a thermal jacket in
order to maintain the optimal temperature and equilibrium for the biological
system. Otherwise, large-scale bioreactors possess an internal coil heating-
cooling system.
Apple Academic Press
On the other hand, as seen before, STRB is one of the most common
systems for lignocellulosic bioconversion and in order to heat transfer and
homogenization be adequate different types of agitation have been studied.
It is worth to mention that agitation promotes the equal distribution and
interaction among all the biological matter and the mixing efficiency will
result in an optimal product yield. As seen in Figure 9.4, there are different
types of impellers that can achieve different mixing patterns as axial or
Author Copy
radial, depending on the movement direction of the components.
Three blade elephant ear impellers (Figure 9.4(A)) are commonly used
in STRB. It has been showed that a configuration of two impellers of this
type have two lowered energy consumption, diminished mixing time and
enhanced homogenization of the medium. At the same time, this configuration
has produced higher glucose conversion by diminishing the inhibition rate of
soluble by-products [70]. Peg mixer (Figure 9.4(B)) has been recognized for
being more convenient on enzymatic saccharification at high solid loadings
where viscosity is challenging, due it promotes the enzymatic binding on the
cellulose substrate and overcomes mass transfer limitations [71].
On the other hand, helical ribbon impellers (Figure 9.4(C)) are used for non-
Newtonian fluids and substrates with high viscosity that present challenges for
the mixing process. This impeller design allows to provide and efficient mixing
not only for this type of fluids but an alternative for enzymatic hydrolysis at
high solid loadings [72]. Besides, it has been demonstrated that this impeller
enhances glucose yields and ethanol production by using less energy [73].
Finally, Rushton turbines are commonly present in most of the vessels
which promotes a radial flow mixing pattern (Figure 9.4(D)). These impellers
have been recently combined with other configurations as the elephant ear
impeller due combines the axial and radial flow overcoming the disadvantage
of Rushton turbines, which creates a staged mixing patters that produces an
irregular enzyme-substrate interaction [70].
The main parameters that must be under tight control and play a crucial
role in the effectiveness of enzymatic catalysis are substrate concentration,
enzyme concentration, temperature, pH, and mixing velocity.
Author Copy
FIGURE 9.4 Different impeller designs for bioreactor systems. (A) Elephant ear impeller;
(B) peg mixer; (C) helical ribbon impeller; and (D) Rushton turbine.
2L-batch reactor Barley straw 15% 7.5 FPU g–1 solids plus 13 CBU g–1 solids 50°C, pH 5, 57 rpm. 81% [126]
–1
250 ml-SSF Barley straw 15% 7 FPU g solids plus 8.4 IU β-glucosidase 50°C, pH 5.5, 150 59% [127]
g–1 solids plus 72 U xylanase g–1 solids rpm.
250 ml batch reactor Wheat straw 20% 7 FPU g–1 DM plus β-glucosidase 50°C, pH 4.8, 6.6 rpm 60% [128]
–1 –1
Batch reactor Rye straw 17% 13 FPU g cellulose plus 35 CBU g 45°C, pH 5, 120 rpm 65% [129]
cellulose
Bagasse
150 ml batch reactor Sugarcane bagasse C. acetobutylicum ATCC 824, 9.6 FPU g–1 50°C, pH 6.7, 120 rpm 55% [130]
Author Copy
250 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
are materials expect a long lifetime [78]. Additionally, the batch system is
the simplest operation mode which implies the least risk of contamination.
Meanwhile, the continuous mode fresh culture medium is continuously added
and extracted from the bioreactor [79]. The volume of the bioreactor is kept
constant because both the inflow and outflow are equal. Normally for this
kind of process is used a membrane for filtration of the ethanol [80]. In this
way is easy to recover a higher amount of ethanol avoiding the inhibition of
the ethanol that limits the growth of the microorganism [81].
Author Copy
FIGURE 9.5 Design of stirred tank bioreactor detailing the process conditions involved in
fermentation process.
outflow where fresh medium is added and part of the inhibitor products is
removed [86].
For the stirring system for the mixing of the substrate, we can identify
several options of mixing from propeller to plates, and as well the baffles
on the reactor walls. Other accessories on the bioreactor that support the
stirring are the air diffusers, but also its combination with mechanical
stirring is another strategy that provides a homogeneous in reaction when
Author Copy
the broth presents resistance to mixing due to a saturation of substrate
or huge growth of microorganism [87]. Stirred systems with successful
results is using a blade stirrer for radial flow, it gives easiness of mixing
from powders to fluids [88, 89]. Meanwhile, the air diffusers are most
used in columns bioreactors in combination with continuous stirred tank
reactors (CSTR).
The measurement of cells in fermentation is usually the cheaper analysis
for the cell counting in Neubauer chamber via Microscopium. However,
the disadvantage is still the no on-line measurement of cells [90, 91]. More
sophisticated instrument is the use of fluorescent in situ hybridization
(FISH) however is expensive. Isaka et al. [92] compared both analyzes, the
cell counting by Neubauer chamber still more versatile for simplicity and
low cost. In new development, several works apply different techniques
to counting cells. One of them is the on-chip as high-speed camera with
high optical resolution and flow technology record to analyze in real-time
the flows [93]. Sobahi and Han [94] fabricated an on-chip system of two
layers; a crystal substrate with electrode detection and the other microfluid
channel layer on top and used Saccharomyces cerevisiae at 0.8×106 cell
mL–1 and 1×106 cell mL–1 concentration diluted in peptone and dextrose
to test.
The temperature measurement plays one of the key roles to keep high
performance in fermentation, moreover, it needs to adjust flows on jackets
and homogeneous stirring in the reactor. Temperature sensors commonly are
added into a middle place of the reactor or close to the bottom lower, but
the instrument sensor does not remain immersed inside the volume reac-
tion, usually are external in the reactor [95]. For lab reactors also is used
an electric heater, Roukas and Kotzekidou [96] used an electric heater in a
batch reactor to adjust to 30°C a fermentation by Saccharomyces cerevisiae.
Author Copy
reactors is normally carried out by passing steam through pipes and a jacket
for a certain time, for the sterilization of substrates the most common is
to sterilize in the same way, passing steam through pipes and a jacket at
a pressure of 0.1 mPa, or sterilize the substrate and solutions in industrial
autoclaves and subsequently being injected through peristaltic pumps into
the reactor, however, the sterilization potential can be reduced if the cleaning
of the reactor is not efficient, if there are sediment residues from previous
processes, the risk of contamination increases dramatically. However, some
accessories cannot be sterilized many times or in the same way, i.e., if the
method used is by moist heat or dry heat sterilization. Although the bioreactor
design allows high temperatures for its sterilization, also with the use of
several times, accessories like pH sensor can be suffer a sensitivity loss. The
alternatives to not apply high temperatures are the use of gamma rays, ozone
gas, or oxygen radicals [99, 100].
Author Copy
tested the viability of microorganism at different ethanol concentration
(v v–1) produced in the fermentation process, by using a concentration of
5×107 cell mL–1 to ferment a broth of 100 g L–1 of reducing sugars and in
a batch reactor system. In comparison with other bioreactor system such
as, fed-batch reactor with the same concentration of cells obtained a high
production of ethanol. Another alternative is the use of pellets with a certain
volume of cells, Ju and Ho [102] reported a 0.66 g g–1 in cell pellet, with a
cell density of 2.57 × 103.
The stoichiometric conversion of glucose to ethanol is 0.51 g g–1. Achieving
this conversion efficiency is difficult because microorganisms use a certain
part of the carbon consumed into cellular metabolism and growth. The yeast
S. cerevisiae is one of the microorganisms that manages to produce ethanol
from high loads of substrate, being able to ferment in media containing up
to 400 g L–1 of glucose. It is important to consider this parameter because
microorganisms can be inhibited by high concentrations of substrate as well
[103]. The concentration of reducing sugars depending on the resistance of
S. cerevisiae, the concentration of ethanol could reach the 11–13% (w v–1)
[104, 105]. Pereira et al. [106] used cellulosic hydrolysates for alcoholic
fermentation and it is carried out in 250 mL Erlenmeyer flasks. They use
anhydrous glucose to reach the concentration of 100 g L–1. These experiments
were conducted for 8 h. Ariyajaroenwong et al. [107] used immobilized cells
of S. cerevisiae in sweet sorghum stalk for the fermentation of sweet sorghum
juice that had a concentration of between 120 and 280 g L–1 of total sugars. In
this case, no agitation was used to prevent the detachment of the immobilized
cells from the carriers.
Several substrates have different nutrient ratios, in the case of the C:N
ratio, the relation of the carbon fermentable of lignocellulose substrates are
interesting in the election for the facility to add less amount of nutrients
Author Copy
romyces cerevisiae [108, 109], although the temperature ratio can vary by
the aim of growth of yeast, Lip et al. [110] reported the growth of yeast
in several temperature conditions, where the maximum growth rate was
between 30–34°C. Most of the fermenting broth used for bioethanol produc-
tion has pH in the range of 4.5–5.5 tested in different sugars concentrations.
The pH could improve de cell growth and ethanol production or can inhibit
the microorganisms in the acid or alkali medium; however, several studies
demonstrated an optimum range for pH conditions between 4–6 [111].
Table 9.4 shows yeast strains used in bioethanol production at different
process conditions of agitation and temperature and as well the concentra-
tion of ethanol obtained from certain sugar concentration.
9.5 CONCLUSIONS
The literature reviewed in this chapter indicates that the field of novel reac-
tors designs and operation process has been intensifying in pretreatment,
enzymatic hydrolysis, and fermentation processes. This has made viable the
use of a great range of LCB and cross the gap of lab-scale to biorefinery
concept. Properly established bioreactor configuration and complex process
conditions provides an important approach insight into the lignocellulosic
bioethanol industry. The main enhancements to scale up biofuels production
process, might be related to develop novel reactors and integrated process
capable to reduce net energy consumption, as well as combined technologies
for better yields in sugars and bioethanol concentration. Finally, tecno-
economical scenarios of the design of new bioreactors configuration must
be considered to evaluate the commercial value of bioethanol production
through LCB since commercial point of view.
TABLE 9.4 Yeast Strains Used in Bioethanol Production at Different Process Conditions
Biomass Inoculated Yeast Cells Sugar Concentration Ethanol Yield T (°C) Agitation References
(g L–1) (%) (rpm)
Sugarcane bagasse 25 g L–1 (dry basis) S. cerevisiae 100 (glucose) 54.69 31 100 [106]
–1
Sugarcane tops 25 g L (dry basis) S. cerevisiae 100 (glucose) 65.52 31 100
Sugarcane straw 25 g L–1 (dry basis) S. cerevisiae 100 (glucose) 74.70 31 100
–1
Sweet sorghum 0.38 g L S. cerevisiae NP 01 228 87.72 30 – [107]
7 –1
Sugarcane bagasse 7.8×10 cell mL S. cerevisiae AR5 78.6 (glucose) 95 33 120 [16]
Spent coffee grounds 1 g/L S. cerevisiae RL-11 52.5 26 30 200 [133]
Sorghum stover S. cerevisiae MTCC 173 200 34 30 120 [134]
Giant reed S. stipitis CBS 6054 33.4 33 30 150 [135]
Corn stover S. cerevisiae ZU-10 66.9 41.6 30 120 [136]
Ipomea carnea S. cerevisiae RPRT90 72.1 46.1 30 150 [137]
Wood S. cerevisiae 37.47 49 30 150 [138]
Reed S. cerevisiae ATCC 123 93 38 150 [139]
Water hyacinth Kluyveromyces marxianus K213 23.3 16 42 – [140]
Wheat straw 1 g/L Kluyveromyces marxianus CECT 22.8 45 42 150 [141]
Author Copy
Design and Engineering Parameters of Bioreactors 257
KEYWORDS
• bioethanol
Apple Academic Press
• bioreactor design
• enzymatic hydrolysis
• fermentation
• pretreatment
• process parameters
Author Copy
REFERENCES
1. Atelge, M. R., Atabani, A. E., Banu, J. R., Krisa, D., Kaya, M., Eskicioglu, C., & Duman,
F., (2020). A critical review of pretreatment technologies to enhance anaerobic digestion
and energy recovery. Fuel, 270, 117494. https://doi.org/10.1016/j.fuel.2020.117494
(accessed on 8 December 2021).
2. Alalwan, H. A., Alminshid, A. H., & Aljaafari, H. A. S., (2019). Promising evolution of
biofuel generations. Subject review. Renewable Energy Focus, 28, 127–139. https://doi.
org/10.1016/j.ref.2018.12.006 (accessed on 8 December 2021).
3. Li, J., Zhou, W., Fan, S., Xiao, Z., Liu, Y., Liu, J., & Wang, Y., (2018). Bioethanol produc-
tion in vacuum membrane distillation bioreactor by permeate fractional condensation
and mechanical vapor compression with polytetrafluoroethylene (PTFE) membrane.
Bioresource Technology, 268, 708–714. https://doi.org/10.1016/j.biortech.2018.08.055
(accessed on 8 December 2021).
4. Toor, M., Kumar, S. S., Malyan, S. K., Bishnoi, N. R., Mathimani, T., Rajendran, K., &
Pugazhendhi, A., (2020). An overview on bioethanol production from lignocellulosic
feedstocks. Chemosphere, 242, 125080. https://doi.org/10.1016/j.chemosphere.2019.
125080 (accessed on 8 December 2021).
5. Sharma, S., Kundu, A., Basu, S., Shetti, N. P., & Aminabhavi, T. M., (2020). Sustainable
environmental management and related biofuel technologies. Journal of Environmental
Management, 273, 111096. https://doi.org/10.1016/j.jenvman.2020.111096 (accessed
on 8 December 2021).
6. Hernández-Beltrán, J. U., Hernández-De, L. I. O., Cruz-Santos, M. M., Saucedo-
Luevanos, A., Hernández-Terán, F., & Balagurusamy, N., (2019). Insight into pretreat-
ment methods of lignocellulosic biomass to increase biogas yield: Current state,
challenges, and opportunities. Applied Sciences, 9, 3721. https://doi.org/10.3390/
app9183721 (accessed on 8 December 2021).
7. Prasad, R. K., Chatterjee, S., Mazumder, P. B., Gupta, S. K., Sharma, S., Vairale, M. G.,
& Gupta, D. K., (2019). Bioethanol production from waste lignocelluloses: A review on
microbial degradation potential. Chemosphere, 231, 588–606. https://doi.org/10.1016/j.
chemosphere.2019.05.142 (accessed on 8 December 2021).
2021).
9. De La Fuente, S. N. M., & Alejo, A. A. M., (2019). Challenges of fermentation
engineering. In: Chávez-González, M. L., Balagurusamy, N., & Aguilar, C. N., (eds.),
Advances in Food Bioproducts and Bioprocessing Technologies (pp. 233–264). CRC
Press. https://doi.org/10.1201/9780429331817-11 (accessed on 8 December 2021).
10. Saha, K., Maharana, A., Sikder, J., Chakraborty, S., Curcio, S., & Drioli, E., (2019).
Continuous production of bioethanol from sugarcane bagasse and downstream purifica-
tion using membrane integrated bioreactor. Catalysis Today, 331, 68–77. https://doi.
org/10.1016/j.cattod.2017.11.031 (accessed on 8 December 2021).
Author Copy
11. Rocha-Meneses, L., Raud, M., Orupõld, K., & Kikas, T., (2019). Potential of bioethanol
production waste for methane recovery. Energy, 173, 133–139. https://doi.org/10.1016/j.
energy.2019.02.073 (accessed on 8 December 2021).
12. Ibarra-Gonzalez, P., & Rong, B. G., (2019). A review of the current state of biofuels
production from lignocellulosic biomass using thermochemical conversion routes.
Chinese Journal of Chemical Engineering, 27, 1523–1535. https://doi.org/10.1016/j.
cjche.2018.09.018 (accessed on 8 December 2021).
13. Alayoubi, R., Mehmood, N., Husson, E., Kouzayha, A., Tabcheh, M., Chaveriat, L.,
& Gosselin, I., (2020). Low temperature ionic liquid pretreatment of lignocellulosic
biomass to enhance bioethanol yield. Renewable Energy, 145, 1808–1816. https://doi.
org/10.1016/j.renene.2019.07.091 (accessed on 8 December 2021).
14. Hernández-Beltrán, J. U., Fontalvo, J., & Hernández-Escoto, H., (2020). Fed-batch
enzymatic hydrolysis of plantain pseudostem to fermentable sugars production and the
impact of particle size at high solids loadings. Biomass Conversion and Biorefinery.
https://doi.org/10.1007/s13399-020-00669-2 (accessed on 8 December 2021).
15. Hernández-Beltrán, J. U., & Hernández-Escoto, H., (2018). Enzymatic hydrolysis of
biomass at high-solids loadings through fed-batch operation. Biomass and Bioenergy,
119. https://doi.org/10.1016/j.biombioe.2018.09.020 (accessed on 8 December 2021).
16. Meléndez-Hernández, P. A., Hernández-Beltrán, J. U., Hernández-Guzmán, A., Morales-
Rodríguez, R., Torres-Guzmán, J. C., & Hernández-Escoto, H., (2019). Comparative of
alkaline hydrogen peroxide pretreatment using NaOH and Ca(OH)2 and their effects
on enzymatic hydrolysis and fermentation steps. Biomass Conversion and Biorefinery.
https://doi.org/10.1007/s13399-019-00574-3 (accessed on 8 December 2021).
17. Kucharska, K., Rybarczyk, P., Hołowacz, I., Łukajtis, R., Glinka, M., & Kamiński,
M., (2018). Pretreatment of lignocellulosic materials as substrates for fermentation
processes. Molecules, 23, 1–32. https://doi.org/10.3390/molecules23112937 (accessed
on 8 December 2021).
18. Gusakov, A. V., Sinitsyn, A. P., & Klyosov, A. A., (1985). Kinetics of the enzymatic
hydrolysis of cellulose: 1. A mathematical model for a batch reactor process. Enzyme
and Microbial Technology, 7, 346–352. https://doi.org/10.1016/0141-0229(85)90114-0
(accessed on 8 December 2021).
19. Goodman, B. A., (2020). Utilization of waste straw and husks from rice production: A
review. Journal of Bioresources and Bioproducts, 5, 143–162. https://doi.org/10.1016/j.
jobab.2020.07.001 (accessed on 8 December 2021).
20. Ponnusamy, V. K., Nguyen, D. D., Dharmaraja, J., Shobana, S., Banu, J. R., Saratale,
R. G., & Kumar, G., (2019). A review on lignin structure, pretreatments, fermentation
reactions and biorefinery potential. Bioresource Technology, 271, 462–472. https://doi.
org/10.1016/j.biortech.2018.09.070 (accessed on 8 December 2021).
Apple Academic Press
21. Wang, H., Pu, Y., Ragauskas, A., & Yang, B., (2019). From lignin to valuable products-
strategies, challenges, and prospects. Bioresource Technology, 271, 449–461. https://
doi.org/10.1016/j.biortech.2018.09.072 (accessed on 8 December 2021).
22. Yoo, C. G., Meng, X., Pu, Y., & Ragauskas, A. J., (2020). The critical role of
lignin in lignocellulosic biomass conversion and recent pretreatment strategies: A
comprehensive review. Bioresource Technology, 301, 122784. https://doi.org/10.1016/j.
biortech.2020.122784 (accessed on 8 December 2021).
23. Sharma, B., Larroche, C., & Dussap, C. G., (2020). Comprehensive assessment of 2G
bioethanol production. Bioresource Technology, 313, 123630. https://doi.org/10.1016/j.
Author Copy
biortech.2020.123630 (accessed on 8 December 2021).
24. Baruah, J., Nath, B. K., Sharma, R., Kumar, S., Deka, R. C., Baruah, D. C., & Kalita,
E., (2018). Recent trends in the pretreatment of lignocellulosic biomass for value-
added products. Frontiers in Energy Research, 6, 1–19. https://doi.org/10.3389/
fenrg.2018.00141 (accessed on 8 December 2021).
25. Kumar, B., Bhardwaj, N., Agrawal, K., Chaturvedi, V., & Verma, P., (2020). Current
perspective on pretreatment technologies using lignocellulosic biomass: An emerging
biorefinery concept. Fuel Processing Technology, 199. https://doi.org/10.1016/j.
fuproc.2019.106244 (accessed on 8 December 2021).
26. Kapoor, M., Semwal, S., Satlewal, A., Christopher, J., Gupta, R. P., Kumar, R., &
Ramakumar, S. S. V., (2019). The impact of particle size of cellulosic residue and solid
loadings on enzymatic hydrolysis with a mass balance. Fuel, 245, 514–520. https://doi.
org/10.1016/j.fuel.2019.02.094 (accessed on 8 December 2021).
27. Shimizu, F. L., Monteiro, P. Q., Ghiraldi, P. H. C., Melati, R. B., Pagnocca, F. C., Souza
De, W., & Brienzo, M., (2018). Acid, alkali and peroxide pretreatments increase the
cellulose accessibility and glucose yield of banana pseudostem. Industrial Crops and
Products, 115, 62–68. https://doi.org/10.1016/j.indcrop.2018.02.024 (accessed on 8
December 2021).
28. Niju, S., Nishanthini, T., & Balajii, M., (2020). Alkaline hydrogen peroxide-pretreated
sugarcane tops for bioethanol production—A process optimization study. Biomass
Conversion and Biorefinery, 10, 149–165. https://doi.org/10.1007/s13399–019–00524-z
(accessed on 8 December 2021).
29. Wang, C., Zhang, X., Liu, Q., Zhang, Q., Chen, L., & Ma, L., (2020). A review of
conversion of lignocellulose biomass to liquid transport fuels by integrated refining
strategies. Fuel Processing Technology, 208, 106485. https://doi.org/10.1016/j.fuproc.
2020.106485 (accessed on 8 December 2021).
30. De Oliveira, A. P. V., Sepulchro, A. G. V., & Polikarpov, I., (2020). Enzymes for
lignocellulosic biomass polysaccharides valorization and production of nanomaterials.
Current Opinion in Green and Sustainable Chemistry, 100397. https://doi.org/10.1016/j.
cogsc.2020.100397 (accessed on 8 December 2021).
31. Maitan-Alfenas, G. P., Visser, E. M., & Guimarães, V. M. (2015). Enzymatic hydrolysis
of lignocellulosic biomass: Converting food waste in valuable products. Current Opinion
in Food Science, 1, 44–49. https://doi.org/10.1016/j.cofs.2014.10.001 (accessed on 8
December 2021).
32. Ruiz, H. A., Rodríguez-Jasso, R. M., Fernandes, B. D., Vicente, A. A., & Teixeira, J. A.,
(2013). Hydrothermal processing, as an alternative for upgrading agriculture residues
and marine biomass according to the biorefinery concept: A review. Renewable and
Sustainable Energy Reviews, 21, 35–51. https://doi.org/10.1016/j.rser.2012.11.069
Apple Academic Press
Author Copy
at low cellulase dosage by fed-batch strategy based on optimized accessory enzymes
and additives. Bioresource Technology, 292, 121993. https://doi.org/10.1016/j.biortech.
2019.121993 (accessed on 8 December 2021).
36. Lorenci, W. A., Dalmas, N. C. J., Porto De, S., Vandenberghe, L., De Carvalho, N. D. P.,
Novak, S. A. C., Letti, L. A. J., & Soccol, C. R., (2020). Lignocellulosic biomass: Acid
and alkaline pretreatments and their effects on biomass recalcitrance - conventional
processing and recent advances. Bioresource Technology, 304, 122848. https://doi.
org/10.1016/j.biortech.2020.122848 (accessed on 8 December 2021).
37. Kumar, D., Juneja, A., & Singh, V., (2018). Fermentation technology to improve
productivity in dry-grind corn process for bioethanol production. Fuel Processing
Technology, 173, 66–74. https://doi.org/10.1016/j.fuproc.2018.01.014 (accessed on 8
December 2021).
38. Germec, M., & Turhan, I., (2018). Ethanol production from acid-pretreated and detoxified
rice straw as sole renewable resource. Biomass Conversion and Biorefinery, 8, 607–619.
https://doi.org/10.1007/s13399-018-0310-1 (accessed on 8 December 2021).
39. Amadi, P. U., & Ifeanacho, M. O., (2016). Impact of changes in fermentation time,
volume of yeast, and mass of plantain pseudo-stem substrate on the simultaneous
saccharification and fermentation potentials of African land snail digestive juice and
yeast. Journal of Genetic Engineering and Biotechnology, 14, 289–297. https://doi.
org/10.1016/j.jgeb.2016.09.002 (accessed on 8 December 2021).
40. Rezania, S., Oryani, B., Cho, J., Talaiekhozani, A., Sabbagh, F., Hashemi, B., &
Mohammadi, A. A., (2020). Different pretreatment technologies of lignocellulosic
biomass for bioethanol production: An overview. Energy, 199, 117457. https://doi.org/
10.1016/j.energy.2020.117457 (accessed on 8 December 2021).
41. Tsapekos, P., Kougias, P. G., Egelund, H., Larsen, U., Pedersen, J., Trénel, P., &
Angelidaki, I., (2017). Mechanical pretreatment at harvesting increases the bioenergy
output from marginal land grasses. Renewable Energy, 111, 914–921. https://doi.org/
10.1016/j.renene.2017.04.061 (accessed on 8 December 2021).
42. Gu, H., An, R., & Bao, J., (2018). Pretreatment refining leads to constant particle size
distribution of lignocellulose biomass in enzymatic hydrolysis. Chemical Engineering
Journal, 352, 198–205. https://doi.org/10.1016/j.cej.2018.06.145 (accessed on 8
December 2021).
43. Aguilar-Reynosa, A., Romaní, A., Ma Rodríguez-Jasso, R., Aguilar, C. N., Garrote, G.,
& Ruiz, H. A., (2017). Microwave heating processing as alternative of pretreatment
Author Copy
approach compared with present practices. Biomass Conversion and Biorefinery, 1.
https://doi.org/10.1007/s13399-020-00958-w (accessed on 8 December 2021).
47. Miliotti, E., Dell’Orco, S., Lotti, G., Rizzo, A. M., Rosi, L., & Chiaramonti, D., (2019).
Lignocellulosic ethanol biorefinery: Valorization of lignin-rich stream through hydro-
thermal liquefaction. Energies, 12, 723. https://doi.org/10.3390/en12040723 (accessed
on 8 December 2021).
48. Terán, H. R., Dionízio, R. M., Sánchez, M. S., Prado, C. A., De Sousa, J. R., Da Silva, S.
S., & Santos, J. C., (2020). Hydrodynamic cavitation-assisted continuous pre-treatment
of sugarcane bagasse for ethanol production: Effects of geometric parameters of the
cavitation device. Ultrasonics Sonochemistry, 63, 104931. https://doi.org/10.1016/j.
ultsonch.2019.104931 (accessed on 8 December 2021).
49. Ramirez-Cadavid, D. A., Kozyuk, O., Lyle, P., & Michel, F. C., (2016). Effects of
hydrodynamic cavitation on dry mill corn ethanol production. Process Biochemistry,
51, 500–508. https://doi.org/10.1016/j.procbio.2016.01.001 (accessed on 8 December
2021).
50. Kim, I., Lee, I., Jeon, S. H., Hwang, T., & Han, J. I., (2015). Hydrodynamic cavitation
as a novel pretreatment approach for bioethanol production from reed. Bioresource
Technology, 192, 335–339. https://doi.org/10.1016/j.biortech.2015.05.038 (accessed on
8 December 2021).
51. Hernández-Guzmán, A., Navarro-Gutiérrez, I. M., Meléndez-Hernández, P. A.,
Hernández-Beltrán, J. U., & Hernández-Escoto, H., (2020). Enhancement of alkaline-
oxidative delignification of wheat straw by semi-batch operation in a stirred tank reactor.
Bioresource Technology, 312, 123589. https://doi.org/10.1016/j.biortech.2020.123589
(accessed on 8 December 2021).
52. Nakashima, K., Ebi, Y., Shibasaki-Kitakawa, N., Soyama, H., & Yonemoto, T.,
(2016). Hydrodynamic cavitation reactor for efficient pretreatment of lignocellulosic
biomass. Industrial and Engineering Chemistry Research, 55, 1866–1871. https://doi.
org/10.1021/acs.iecr.5b04375 (accessed on 8 December 2021).
53. Terán, H. R., Ramos, L., Da Silva, S. S., Dragone, G., Mussatto, S. I., & Santos, J.
C. D., (2018). Hydrodynamic cavitation as a strategy to enhance the efficiency of
lignocellulosic biomass pretreatment. Critical Reviews in Biotechnology, 38, 483–493.
https://doi.org/10.1080/07388551.2017.1369932 (accessed on 8 December 2021).
54. Andrić, P., Meyer, A. S., Jensen, P. A., & Dam-Johansen, K., (2010). Reactor design
for minimizing product inhibition during enzymatic lignocellulose hydrolysis. II.
Author Copy
org/10.1002/elsc.201600102 (accessed on 8 December 2021).
58. Du, J., Zhang, F., Li, Y., Zhang, H., Liang, J., Zheng, H., & Huang, H., (2014). Enzymatic
liquefaction and saccharification of pretreated corn stover at high-solids concentrations
in a horizontal rotating bioreactor. Bioprocess and Biosystems Engineering, 37, 173–181.
https://doi.org/10.1007/s00449-013-0983-6 (accessed on 8 December 2021).
59. Hodge, D. B., Karim, M. N., Schell, D. J., & McMillan, J. D., (2009). Model-based
fed-batch for high-solids enzymatic cellulose hydrolysis. Applied Biochemistry and
Biotechnology, 152, 88–107. https://doi.org/10.1007/s12010-008-8217-0 (accessed on
8 December 2021).
60. Al-Zuhair, S., Al-Hosany, M., Zooba, Y., Al-Hammadi, A., & Al-Kaabi, S., (2013).
Development of a membrane bioreactor for enzymatic hydrolysis of cellulose. Renew-
able Energy, 56, 85–89. https://doi.org/10.1016/j.renene.2012.09.044 (accessed on 8
December 2021).
61. Garcia-Ochoa, F., Santos, V. E., & Gomez, E., (2019). Stirred tank bioreactors. Compre-
hensive Biotechnology (3rd edn., Vol. 2). Elsevier. https://doi.org/10.1016/B978-0-444-
64046-8.00078-1 (accessed on 8 December 2021).
62. Burman, N. W., Sheridan, C. M., & Harding, K. G., (2019). Lignocellulosic bioethanol
production from grasses pre-treated with acid mine drainage: Modeling and comparison
of SHF and SSF. Bioresource Technology Reports, 7, 100299. https://doi.org/10.1016/j.
biteb.2019.100299 (accessed on 8 December 2021).
63. Maslova, O., Stepanov, N., Senko, O., & Efremenko, E., (2019). Production of various
organic acids from different renewable sources by immobilized cells in the regimes
of separate hydrolysis and fermentation (SHF) and simultaneous saccharification
and fermentation (SFF). Bioresource Technology, 272, 1–9. https://doi.org/10.1016/j.
biortech.2018.09.143 (accessed on 8 December 2021).
64. Loaces, I., Schein, S., & Noya, F., (2017). Ethanol production by Escherichia coli
from Arundo donax biomass under SSF, SHF or CBP process configurations and in
situ production of a multifunctional glucanase and xylanase. Bioresource Technology,
224, 307–313. https://doi.org/10.1016/j.biortech.2016.10.075 (accessed on 8 December
2021).
65. Mithra, M. G., Jeeva, M. L., Sajeev, M. S., & Padmaja, G., (2018). Comparison of
ethanol yield from pretreated lignocellulose-starch biomass under fed-batch SHF or SSF
modes. Heliyon, 4, e00885. https://doi.org/10.1016/j.heliyon.2018.e00885 (accessed on
8 December 2021).
66. Jin, M., Lau, M. W., Balan, V., & Dale, B. E., (2010). Two-step SSCF to convert AFEX-
treated switchgrass to ethanol using commercial enzymes and Saccharomyces cerevisiae
424A(LNH-ST). Bioresource Technology, 101, 8171–8178. https://doi.org/10.1016/j.
biortech.2010.06.026 (accessed on 8 December 2021).
Apple Academic Press
67. Liu, Z. H., & Chen, H. Z., (2016). Simultaneous saccharification and co-fermentation
for improving the xylose utilization of steam-exploded corn stover at high solid loading.
Bioresource Technology, 201, 15–26. https://doi.org/10.1016/j.biortech.2015.11.023
(accessed on 8 December 2021).
68. M’Barek, H. N., Arif, S., Taidi, B., & Hajjaj, H., (2020). Consolidated bioethanol
production from olive mill waste: Wood-decay fungi from central Morocco as promising
decomposition and fermentation biocatalysts. Biotechnology Reports. https://doi.
org/10.1016/j.btre.2020.e00541 (accessed on 8 December 2021).
69. Rastogi, M., & Shrivastava, S., (2017). Recent advances in second-generation bioethanol
Author Copy
production: An insight to pretreatment, saccharification and fermentation processes.
Renewable and Sustainable Energy Reviews, 80, 330–340. https://doi.org/10.1016/j.
rser.2017.05.225 (accessed on 8 December 2021).
70. Correâ, L. J., Badino, A. C., & Cruz, A. J. G., (2016). Mixing design for enzymatic
hydrolysis of sugarcane bagasse: Methodology for selection of impeller configuration.
Bioprocess and Biosystems Engineering, 39, 285–294. https://doi.org/10.1007/s00449-
015-1512-6 (accessed on 8 December 2021).
71. Caspeta, L., Caro-Bermúdez, M. A., Ponce-Noyola, T., & Martinez, A., (2014).
Enzymatic hydrolysis at high-solids loadings for the conversion of agave bagasse to fuel
ethanol. Applied Energy, 113, 277–286. https://doi.org/10.1016/j.apenergy.2013.07.036
(accessed on 8 December 2021).
72. Modenbach, A. A., & Nokes, S. E., (2013). Enzymatic hydrolysis of biomass at high-
solids loadings - A review. Biomass and Bioenergy, 56, 526–544. https://doi.org/10.1016/
j.biombioe.2013.05.031 (accessed on 8 December 2021).
73. Zhang, J., Chu, D., Huang, J., Yu, Z., Dai, G., & Bao, J., (2010). Simultaneous sacchari-
fication and ethanol fermentation at high corn stover solids loading in a helical stirring
bioreactor. Biotechnology and Bioengineering, 105, 718–728. https://doi.org/10.1002/
bit.22593 (accessed on 8 December 2021).
74. Várnai, A., Siika-Aho, M., & Viikari, L., (2013). Carbohydrate-binding modules (CBMs)
revisited: Reduced amount of water counterbalances the need for CBMs. Biotechnology for
Biofuels, 6, 30. https://doi.org/10.1186/1754-6834-6-30 (accessed on 8 December 2021).
75. Michelin, M., Ruiz, H. A., Silva, D. P., Ruzene, D. S., Teixeira, J. A., & M, M. L. T.,
(2014). Cellulose from lignocellulosic waste. In: Ramawat, K. G., (ed.), Polysaccharides.
Springer International Publishing Switzerland. https://doi.org/10.1007/978-3-319-
03751-6 (accessed on 8 December 2021).
76. Saini, J. K., Patel, A. K., Adsul, M., & Singhania, R. R., (2016). Cellulase adsorption
on lignin: A roadblock for economic hydrolysis of biomass. Renewable Energy, 98,
29–42. https://doi.org/10.1016/j.renene.2016.03.089 (accessed on 8 December 2021).
77. Wood, B. E., Aldrich, H. C., & Ingram, L. O., (1997). Ultrasound stimulates ethanol
production during the simultaneous saccharification and fermentation of mixed waste
office paper. Biotechnology Progress, 13, 232–237. https://doi.org/10.1021/bp970027v
(accessed on 8 December 2021).
78. Ketchum, L. H., (1997). Design and physical features of sequencing batch reactors. Water
Science and Technology, 35(1), 11–18. https://doi.org/10.1016/S0273-1223(96)00873-6
(accessed on 8 December 2021).
79. Kropp, C., Massai, D., & Zweigerdt, R., (2017). Progress and challenges in large-scale
expansion of human pluripotent stem cells. Process Biochemistry, 59, 244–254. https://
doi.org/10.1016/j.procbio.2016.09.032 (accessed on 8 December 2021).
80. Ylitervo, P., Franzén, C. J., & Taherzadeh, M. J., (2014). Continuous ethanol production
Apple Academic Press
Author Copy
(ed.), Alcohol Fuels: Current Technologies and Future Prospect. Intechopen. https://doi.
org/10.5772/intechopen.86701 (accessed on 8 December 2021).
84. Liu, C. G., Xiao, Y., Xia, X. X., Zhao, X. Q., Peng, L., Srinophakun, P., & Bai, F. W.,
(2019). Cellulosic ethanol production: Progress, challenges and strategies for solutions.
Biotechnology Advances, 37, 491–504. https://doi.org/10.1016/j.biotechadv.2019.03.002
(accessed on 8 December 2021).
85. Leuteritz, A., Döring, K. D., Lampke, T., & Kuehnert, I., (2016). Accelerated ageing
of plastic jacket pipes for district heating. Polymer Testing, 51, 142–147. https://doi.
org/10.1016/j.polymertesting.2016.03.012 (accessed on 8 December 2021).
86. Pumphrey, B., & Julien, C., (1996). An introduction to fermentation: Fermentation
basics. Journal of the American Society of Brewing Chemists, 34, 103–104. https://doi.
org/10.1080/03610470.1976.12006198 (accessed on 8 December 2021).
87. Jin, B., & Lant, P., (2004). Flow regime, hydrodynamics, floc size distribution and
sludge properties in activated sludge bubble column, air-lift and aerated stirred reactors.
Chemical Engineering Science, 59(12), 2379–2388. https://doi.org/10.1016/j.ces.2004.
01.061 (accessed on 8 December 2021).
88. Ameur, H., (2016). Mixing of complex fluids with flat and pitched bladed impellers:
Effect of blade attack angle and shear-thinning behavior. Food and Bioproducts
Processing, 99, 71–77. https://doi.org/10.1016/j.fbp.2016.04.004 (accessed on 8
December 2021).
89. Stewart, R. L., Bridgwater, J., & Parker, D. J., (2001). Granular flow over a flat-bladed
stirrer. Chemical Engineering Science, 56(14), 4257–4271. https://doi.org/10.1016/
S0009-2509(01)00104-X (accessed on 8 December 2021).
90. Caligiore-Gei, P. F., & Valdez, J. G. (2015). Adjustment of a rapid method for
quantification of Fusarium spp. spore suspensions in plant pathology. Revista Argentina
de Microbiologia, 47(2), 152–154. https://doi.org/10.1016/j.ram.2015.03.002 (accessed
on 8 December 2021).
91. De Souza, L. A. T., Da Silva, F. E. A., De Morais, J. O. F., Simões, D. A., & De Morais,
M. A., (2005). Contaminant yeast detection in industrial ethanol fermentation must by
rDNA-PCR. Letters in Applied Microbiology, 40, 19–23. https://doi.org/10.1111/j.1472-
765X.2004.01618.x (accessed on 8 December 2021).
92. Isaka, K., Date, Y., Sumino, T., Yoshie, S., & Tsuneda, S., (2006). Growth characteristic
of anaerobic ammonium-oxidizing bacteria in an anaerobic biological filtrated reactor.
Applied Microbiology and Biotechnology, 70, 47–52. https://doi.org/10.1007/s00253-
005-0046-2 (accessed on 8 December 2021).
93. Goda, K., Ayazi, A., Gossett, D. R., Sadasivam, J., Lonappan, C. K., Sollier, E., & Jalali,
B., (2012). High-throughput single-microparticle imaging flow analyzer. Proceedings of
the National Academy of Sciences of the United States of America, 109, 11630–11635.
https://doi.org/10.1073/pnas.1204718109 (accessed on 8 December 2021).
Apple Academic Press
94. Sobahi, N., & Han, A., (2020). High-throughput and label-free multi-outlet cell counting
using a single pair of impedance electrodes. Biosensors and Bioelectronics, 166, 112458.
https://doi.org/10.1016/j.bios.2020.112458 (accessed on 8 December 2021).
95. Darnoko, D., & Cheryan, M., (2000). Kinetics of palm oil transesterification in a batch
reactor. JAOCS, Journal of the American Oil Chemists’ Society, 77, 1263–1267. https://
doi.org/10.1007/s11746-000-0198-y (accessed on 8 December 2021).
96. Roukas, T., & Kotzekidou, P., (2020). Rotary biofilm reactor: A new tool for long-term
bioethanol production from non-sterilized beet molasses by Saccharomyces cerevisiae
in repeated-batch fermentation. Journal of Cleaner Production, 257, 120519. https://
Author Copy
doi.org/10.1016/j.jclepro.2020.120519 (accessed on 8 December 2021).
97. Claros, J., Serralta, J., Seco, A., Ferrer, J., & Aguado, D., (2012). Real-time control
strategy for nitrogen removal via nitrite in a SHARON reactor using pH and ORP sensors.
Process Biochemistry, 47, 1510–1515. https://doi.org/10.1016/j.procbio.2012.05.020
(accessed on 8 December 2021).
98. De Vleeschauwer, F., Caluwé, M., Dobbeleers, T., Stes, H., Dockx, L., Kiekens, F.,
& Dries, J., (2020). A dynamic control system for aerobic granular sludge reactors
treating high COD/P wastewater, using pH and DO sensors. Journal of Water Process
Engineering, 33, 101065. https://doi.org/10.1016/j.jwpe.2019.101065 (accessed on 8
December 2021).
99. Nagatsu, M., Terashita, F., Nonaka, H., Xu, L., Nagata, T., & Koide, Y., (2005). Effects of
oxygen radicals in low-pressure surface-wave plasma on sterilization. Applied Physics
Letters, 86(21), 1–3. https://doi.org/10.1063/1.1931050 (accessed on 8 December 2021).
100. Takatsuji, Y., Ishikawa, S., & Haruyama, T., (2017). Efficient sterilization using reactive
oxygen species generated by a radical vapor reactor. Process Biochemistry, 54, 140–143.
https://doi.org/10.1016/j.procbio.2017.01.002 (accessed on 8 December 2021).
101. Godbey, W. T., (2014). Fermentation, Beer, and Biofuels. An Introduction to Biotech-
nology. Elsevier Ltd. https://doi.org/10.1016/b978-1-907568-28-2.00016-2 (accessed
on 8 December 2021).
102. Ju, L. K., & Ho, C. S., (1988). Correlation of cell volume fractions with cell
concentrations in fermentation media. Biotechnology and Bioengineering, 32, 95–99.
https://doi.org/10.1002/bit.260320113 (accessed on 8 December 2021).
103. Moreno, A. D., Alvira, P., Ibarra, D., & Tomás-Pejó, E. (2017). Production of ethanol
from lignocellulosic biomass. In Z. Fang, R. L. Smith, & J. X. Qi (Eds.), Production
of Platform Chemicals from Sustainable Resources (pp. 375–410). Springer Nature
Singapore Pte Ltd. https://doi.org/10.1007/978-981-10-4172-3_12 (accessed on 8
December 2021).
104. Ghareib, M., Youssef, K. A., & Khalil, A. A., (1988). Ethanol tolerance of Saccharomyces
cerevisiae and its relationship to lipid content and composition. Folia Microbiologica,
33(6), 447–452. https://doi.org/10.1007/BF02925769 (accessed on 8 December 2021).
105. Stanley, D., Bandara, A., Fraser, S., Chambers, P. J., & Stanley, G. A., (2010). The
ethanol stress response and ethanol tolerance of Saccharomyces cerevisiae. Journal of
Applied Microbiology, 109, 13–24. https://doi.org/10.1111/j.1365-2672.2009.04657.x
(accessed on 8 December 2021).
106. Pereira, S. C., Maehara, L., Machado, C. M. M., & Farinas, C. S., (2015). 2G ethanol
from the whole sugarcane lignocellulosic biomass. Biotechnology for Biofuels, 8, 1–16.
https://doi.org/10.1186/s13068-015-0224-0 (accessed on 8 December 2021).
107. Ariyajaroenwong, P., Laopaiboon, P., Salakkam, A., Srinophakun, P., & Laopaiboon,
Apple Academic Press
L., (2016). Kinetic models for batch and continuous ethanol fermentation from sweet
sorghum juice by yeast immobilized on sweet sorghum stalks. Journal of the Taiwan
Institute of Chemical Engineers, 66, 210–216. https://doi.org/10.1016/j.jtice.2016.06.023
(accessed on 8 December 2021).
108. De Souza, D. M. O., Maciel, F. R., Mantelatto, P. E., Cavalett, O., Rossell, C. E. V.,
Bonomi, A., & Leal, M. R. L. V., (2015). Sugarcane processing for ethanol and
sugar in Brazil. Environmental Development, 15, 35–51. https://doi.org/10.1016/j.
envdev.2015.03.004 (accessed on 8 December 2021).
109. Veloso, I. I. K., Rodrigues, K. C. S., Sonego, J. L. S., Cruz, A. J. G., & Badino, A. C.,
Author Copy
(2019). Fed-batch ethanol fermentation at low temperature as a way to obtain highly
concentrated alcoholic wines: Modeling and optimization. Biochemical Engineering
Journal, 141, 60–70. https://doi.org/10.1016/j.bej.2018.10.005 (accessed on 8 December
2021).
110. Lip, K. Y. F., García-Ríos, E., Costa, C. E., Guillamón, J. M., Domingues, L., Teixeira,
J., & Van, G. W. M., (2020). Selection and subsequent physiological characterization of
industrial Saccharomyces cerevisiae strains during continuous growth at sub- and- supra
optimal temperatures. Biotechnology Reports, 26. https://doi.org/10.1016/j.btre.2020.
e00462 (accessed on 8 December 2021).
111. Izmirlioglu, G., & Demirci, A., (2010). Ethanol production from waste potato mash
by using Saccharomyces cerevisiae. American Society of Agricultural and Biological
Engineers Annual International Meeting 2010, 2, 1571–1581. https://doi.org/10.3390/
app2040738 (accessed on 8 December 2021).
112. Lin, W., Xing, S., Jin, Y., Lu, X., Huang, C., & Yong, Q. (2020). Insight into under-
standing the performance of deep eutectic solvent pretreatment on improving enzymatic
digestibility of bamboo residues. Bioresource Technology, 306, 123163. https://doi.
org/10.1016/j.biortech.2020.123163 (accessed on 8 December 2021).
113. Tan, L., Zhong, J., Jin, Y. L., Sun, Z. Y., Tang, Y. Q., & Kida, K. (2020). Production of
bioethanol from unwashed-pretreated rapeseed straw at high solid loading. Bioresource
Technology, 303, 122949. https://doi.org/10.1016/j.biortech.2020.122949 (accessed on
8 December 2021).
114. Kainthola, J., Shariq, M., Kalamdhad, A. S., & Goud, V. V. (2019). Enhanced methane
potential of rice straw with microwave assisted pretreatment and its kinetic analysis.
Journal of Environmental Management, 232, 188–196. https://doi.org/10.1016/j.jenvman.
2018.11.052 (accessed on 8 December 2021).
115. Shen, J., Zheng, Q., Zhang, R., Chen, C., & Liu, G. (2019). Co-pretreatment of wheat
straw by potassium hydroxide and calcium hydroxide: Methane production, economics,
and energy potential analysis. Journal of Environmental Management, 236, 720–726.
https://doi.org/10.1016/j.jenvman.2019.01.046 (accessed on 8 December 2021).
116. Pérez-Pimienta, J. A., Papa, G., Gladden, J. M., Simmons, B. A., & Sanchez, A.
(2020). The effect of continuous tubular reactor technologies on the pretreatment of
lignocellulosic biomass at pilot-scale for bioethanol production. RSC Advances, 10,
18147–18159. https://doi.org/10.1039/d0ra04031b (accessed on 8 December 2021).
117. Niglio, S., Procentese, A., Russo, M. E., Piscitelli, A., & Marzocchella, A. (2019).
Integrated enzymatic pretreatment and hydrolysis of apple pomace in a bubble column
bioreactor. Biochemical Engineering Journal, 150, 107306. https://doi.org/10.1016/j.
bej.2019.107306 (accessed on 8 December 2021).
Apple Academic Press
118. Miranda, F. S., Rabelo, S. C., Pradella, J. G. C., Carli, C. Di, Petraconi, G., Maciel, H.
S., … Vieira, L. (2020). Plasma in-Liquid Using Non-contact Electrodes: A Method
of Pretreatment to Enhance the Enzymatic Hydrolysis of Biomass. Waste and Biomass
Valorization, 11, 4921–4931. https://doi.org/10.1007/s12649-019-00824-5 (accessed on
8 December 2021).
119. Zeng, G., You, H., Wang, K., Jiang, Y., Bao, H., Du, M., … Gu, Z. (2020). Semi-
simultaneous Saccharification and Fermentation of Ethanol Production from Sargassum
horneri and Biosorbent Production from Fermentation Residues. Waste and Biomass
Valorization, 11, 4743–4755. https://doi.org/10.1007/s12649-019-00748-0 (accessed on
Author Copy
8 December 2021).
120. de Carvalho Silvello, M. A., Martínez, J., & Goldbeck, R. (2020). Application of
Supercritical CO2 Treatment Enhances Enzymatic Hydrolysis of Sugarcane Bagasse.
Bioenergy Research, 13, 786–796. https://doi.org/10.1007/s12155-020-10130-x (accessed
on 8 December 2021).
121. Yuan, Z., Wei, W., Wen, Y., & Wang, R. (2019). Comparison of alkaline and acid-
catalyzed steam pretreatments for ethanol production from tobacco stalk. Industrial
Crops and Products, 142, 111864. https://doi.org/10.1016/j.indcrop.2019.111864
(accessed on 8 December 2021).
122. Antunes, F. A. F., Chandel, A. K., Brumano, L. P., Terán Hilares, R., Peres, G. F. D.,
Ayabe, L. E. S., … Da Silva, S. S. (2018). A novel process intensification strategy
for second-generation ethanol production from sugarcane bagasse in fluidized bed
reactor. Renewable Energy, 124, 189–196. https://doi.org/10.1016/j.renene.2017.06.004
(accessed on 8 December 2021).
123. Matsakas, L., Nitsos, C., Raghavendran, V., Yakimenko, O., Persson, G., Olsson, E., …
Christakopoulos, P. (2018). A novel hybrid organosolv: Steam explosion method for the
efficient fractionation and pretreatment of birch biomass. Biotechnology for Biofuels, 11,
160. https://doi.org/10.1186/s13068-018-1163-3 (accessed on 8 December 2021).
124. Yang, J., Zhang, X., Yong, Q., & Yu, S. (2011). Three-stage enzymatic hydrolysis of
steam-exploded corn stover at high substrate concentration. Bioresource Technology,
102, 4905–4908. https://doi.org/10.1016/j.biortech.2010.12.047 (accessed on 8 December
2021).
125. Lau, M. W., Dale, B. E., & Balan, V. (2008). Ethanolic fermentation of hydrolysates
from ammonia fiber expansion (AFEX) treated corn stover and distillers grain without
detoxification and external nutrient supplementation. Biotechnology and Bioengineering,
99, 529–539. https://doi.org/10.1002/bit.21609 (accessed on 8 December 2021).
126. Rosgaard, L., Andric, P., Dam-Johansen, K., Pedersen, S., & Meyer, A. S. (2007).
Effects of substrate loading on enzymatic hydrolysis and viscosity of pretreated barley
straw. Applied Biochemistry and Biotechnology, 143, 27–40. https://doi.org/10.1007/
s12010-007-0028-1 (accessed on 8 December 2021).
127. García-Aparicio, M. P., Oliva, J. M., Manzanares, P., Ballesteros, M., Ballesteros, I.,
González, A., & Negro, M. J. (2011). Second-generation ethanol production from steam
exploded barley straw by Kluyveromyces marxianus CECT 10875. Fuel, 90, 1624–1630.
128. Jørgensen, H., Vibe-Pedersen, J., Larsen, J., & Felby, C. (2006). Liquefaction of ligno-
cellulose at high-solids concentrations. Biotechnology and Bioengineering, 96, 862–870.
https://doi.org/10.1002/bit.21115 (accessed on 8 December 2021).
129. Ingram, T., Wörmeyer, K., Lima, J. C. I., Bockemühl, V., Antranikian, G., Brunner, G.,
Apple Academic Press
Author Copy
consistency enzymatic saccharification of sweet sorghum bagasse pretreated with
liquid hot water. Bioresource Technology, 108, 252–257. https://doi.org/10.1016/j.
biortech.2011.12.092 (accessed on 8 December 2021).
132. Phukoetphim, N., Salakkam, A., Laopaiboon, P., & Laopaiboon, L. (2017). Improvement
of ethanol production from sweet sorghum juice under batch and fed-batch fermentations:
Effects of sugar levels, nitrogen supplementation, and feeding regimes. Electronic Journal
of Biotechnology, 26, 84–92. https://doi.org/10.1016/j.ejbt.2017.01.005 (accessed on 8
December 2021).
133. Mussatto, S. I., Machado, E. M. S., Carneiro, L. M., & Teixeira, J. A. (2012). Sugars
metabolism and ethanol production by different yeast strains from coffee industry wastes
hydrolysates. Applied Energy, 92, 763–768. https://doi.org/10.1016/j.apenergy.2011.
08.020 (accessed on 8 December 2021).
134. Sathesh-Prabu, C., & Murugesan, A. G. (2011). Potential utilization of sorghum field
waste for fuel ethanol production employing Pachysolen tannophilus and Saccharomyces
cerevisiae. Bioresource Technology, 102(3), 2788–2792. https://doi.org/10.1016/j.biortech.
2010.11.097 (accessed on 8 December 2021).
135. Scordia, D., Cosentino, S. L., Lee, J. W., & Jeffries, T. W. (2012). Bioconversion of giant
reed (Arundo donax L.) hemicellulose hydrolysate to ethanol by Scheffersomyces stipitis
CBS6054. Biomass and Bioenergy, 39, 296–305. https://doi.org/10.1016/j.biombioe.
2012.01.023 (accessed on 8 December 2021).
136. Zhao, J., & Xia, L. (2010). Bioconversion of corn stover hydrolysate to ethanol by a
recombinant yeast strain. Fuel Processing Technology, 91(12), 1807–1811. https://doi.
org/10.1016/j.fuproc.2010.08.002 (accessed on 8 December 2021).
137. Kumari, R., & Pramanik, K. (2013). Bioethanol production from Ipomoea Carnea biomass
using a potential hybrid yeast strain. Applied Biochemistry and Biotechnology, 171(3),
771–785. https://doi.org/10.1007/s12010-013-0398-5 (accessed on 8 December 2021).
138. Gupta, R., Sharma, K. K., & Kuhad, R. C. (2009). Separate hydrolysis and fermentation
(SHF) of Prosopis juliflora, a woody substrate, for the production of cellulosic ethanol by
Saccharomyces cerevisiae and Pichia stipitis-NCIM 3498. Bioresource Technology, 100(3),
1214–1220. https://doi.org/10.1016/j.biortech.2008.08.033 (accessed on 8 December 2021).
139. Li, H., Kim, N. J., Jiang, M., Kang, J. W., & Chang, H. N. (2009). Simultaneous
saccharification and fermentation of lignocellulosic residues pretreated with phosphoric
acid-acetone for bioethanol production. Bioresource Technology, 100(13), 3245–3251.
https://doi.org/10.1016/j.biortech.2009.01.021 (accessed on 8 December 2021).
140. Yan, J., Wei, Z., Wang, Q., He, M., Li, S., & Irbis, C. (2015). Bioethanol production
from sodium hydroxide/hydrogen peroxide-pretreated water hyacinth via simultaneous
saccharification and fermentation with a newly isolated thermotolerant Kluyveromyces
marxianu strain. Bioresource Technology, 193, 103–109. https://doi.org/10.1016/j.
Apple Academic Press
Author Copy
2021).
Author Copy
P. ABDESHAHIAN,1 S. S. DA SILVA,1 N. BALAGURUSAMY,5 and
J. C. SANTOS2
Biopolymers, Bioprocesses, Process Simulation Laboratory,
1
ABSTRACT
the use of corn in the off-season of sugarcane, thus increasing the annual
production of bioethanol. In this concept, there is also the possibility of flex
biorefinery with 1G and 2G ethanol production. In fact, the conversion of
sugarcane biomass into the fermentable sugars for the production of second-
generation ethanol is a promising alternative to meet the future demands
for this biofuel and could be successfully integrated with sugarcane and
corn-based 1G facilities. However, in order to attain such success and to take
Author Copy
advantage of those flexible industries, the construction of a bridge between
science and industry is essential. In this view, investments in research and
development with the transfer of new technologies from the academy to the
industry are required. Furthermore, the training of skilled labor to deal with
new technological challenges is essential.
10.1 INTRODUCTION
Currently, there are many projects for the production of bioethanol across
the world, which is known greatly in line with the corn biorefinery in USA
and sugarcane biorefinery in Brazil [176], while there are other countries
investing in the utilization of this biofuel. For example, in America, Mexico
Apple Academic Press
Author Copy
The global ethanol production in 2019 was 37.8 billion gallons in which
USA and Brazil had 83% and 26% of the total, respectively [183]. In the
United States, corn is the main raw material to produce ethanol so that in
2017 it represented 10% of the demand for vehicular fuel. In Brazil, on
the other hand, sugarcane (juice or molasses) is the main raw material for
ethanol production [176].
Raw material choice is the fundamental step for the ethanol industry,
which can represent up to 42% of the production costs [184, 185]. In addi-
tion, the commercial adoption of biofuels depends on the evaluation of the
broad levels of efficiency [185, 186] considering several economic, social,
environmental, and strategic criteria, such as national energy security. Hence,
in order to use a source of bioenergy, the chosen biomass needs to meet
these requirements. The ideal characteristics for using a biomass feedstock
as a source of energy include high agricultural production, favorable natural
cycles, low energy consumption in its cultivation, a low production cost, the
low levels of contaminants, and a low demand for nutrients [6]. Furthermore,
it is important that the selection of biomass can favor the carbon balance
when considering life cycle of the biofuel, thus it can take into account the
entire production and use in the production chain [175, 176, 184, 185].
Currently, ethanol can be classified as the first generation (1G) when it is
produced from food crops such as sucrose and starch. Ethanol is known as
the second generation (2G) when it is produced from biomass residues and
byproducts such as sugarcane bagasse and corn cob, while third-generation
(3G) of ethanol is referred to the ethanol produced from microbial biomass,
namely microalgae. In this context, some authors indicate ethanol as the
fourth generation (4G) of ethanol when it is produced from genetically
modified organism, e.g., cyanobacteria through a process named ‘photo-
fermentation’ (direct conversion of light and carbon dioxide (CO2) into
ethanol) [7]. Moreover, ethanol has been called as 1.5G when it is produced
in the 1G industry, but additional sugars from cellulosic fibers is utilized, for
example, from corn grain [8].
These concepts are important to a systematic study of new options for
Apple Academic Press
the integration of biorefineries and raw materials. In Brazil, the main raw
material for ethanol production is sugarcane juice that is produced only
for the alcohol, or as more commonly observed, sugarcane molasses used
in integrated sugar and ethanol industries [9]. However, Brazil is a great
producer of corn, mainly in some regions such as central west of the country.
Thus, sugarcane has an interseason period from December to April, which is
an interesting opportunity for 1G biorefineries using sugarcane with corn as
Author Copy
raw materials in flex biorefineries [184].
Indeed, corn and sugarcane biorefineries are established processes
used in the world to produce ethanol [3, 4, 10] such as Brazil so that such
countries could integrate those raw materials and take advantage of their
characteristics. The aim of the present chapter is to discuss the possibility of
the integration of the use of both raw materials (corn and sugarcane) and to
give future perspectives of the integration in flex biorefineries. In this way, a
brief general overview is presented in the first sections.
residues.
TABLE 10.1 Main World Carbon Sources (Starchy and Sucrose-based Ones) for Ethanol
Production
Countries Biomass Ethanol Production Biomass Residual References
Production (million gallons) in Biorefinery
(million tons) (million tons)
Author Copy
United States 342 (corn) 149 75 (corn ethanol 68.25 (corn straw) [13, 14]
(data of 2018) (biomass from production)
corn for ethanol
production)
Brazil (data of 642.7 26 (sugarcane ethanol 96.3 (sugarcane [11, 15,
2019) (sugarcane) production) 2.5 (corn bagasse 189]
95.60 (corn) ethanol production lignocellulosic)
Europe Union 61.60 (corn) 13 (corn) 21 (beet) 64.28 (wheat [7, 14]
(data of 2019) 168.28 (wheat) straw)
35.82 (beet)
China (data of 215.00 (corn) 8.45 (corn) – [16]
2015)
Canada (data 203 (corn) 4.36 (corn) – [17]
of 2019)
India (data of 115.6 (rice) 2.23 (rice) 22.57 (rice straw) [18, 190]
2019) 85.72 (corn)
21.16 (wheat)
the interseason period of sugarcane cultivation in Brazil [19, 192]. Thus, the
utilization of sugarcane and corn reveals an interesting example of a highly
flexible and productive biorefinery.
Apple Academic Press
10.2.1 CORN
The raw materials used for the first-generation fuel ethanol are mainly crops
rich in starch or sucrose such as corn, cassava, and sugarcane [2, 20]. Corn
(Zea mays) is a cereal grain, which is widely grown all over the world. It is
widely used as a staple food for humans or as animal feed, due to its numerous
Author Copy
nutritional properties. All studies suggest that its origin is Mexican since its
domestication started 7,500 to 12,000 years ago in the center of Mexico [21].
Maize contains facilities for its cultivation and therefore has a great potential
for production and adaptation to technology. It has a mechanized cultivation
that benefits greatly from modern planting and harvesting techniques [20].
The world production of corn was 850 million tons in 2019, which was more
than rice (678 million tons) and wheat (682 million tons) [184]. Corn is
grown in different regions of the world. The largest producers are the USA
and China, which have produced 37 and 21% of the total world production,
respectively [11, 22].
10.2.2 SUGARCANE
The sugarcane cycle plant has an average of six years, in which four or five
harvests occur. Sugarcane belongs to a group of tall perennial grass species
of the genus Saccharum. This plant is native in the tropical regions such as
the South of Asia and North of Brazil, and can be used for the production
of ethanol and sugar [193]. Its stems are robust, articulated, and fibrous.
Sugarcane height can be two to six meters. Researchers have been mixed
sugarcane species to develop complex hybrid for enhancing its performance
in different areas. Sugarcane belongs to economically important plants such
as corn, wheat, rice, and sorghum [194, 195]. Sucrose is the principal sugar
obtained and can be extracted and purified in biorefineries for further use as a
raw material in the food industry, or it would be fermented to produce ethanol
[196]. In 2019, sugarcane grew on about 26.0 million hectares of agricultural
lands in more than 90 countries, with a worldwide harvest of 174 million tons
[184]. Brazil is the largest producer of sugarcane in the world. Other largest
producers are Thailand, India, China, Pakistan, and Mexico [11].
Author Copy
works and what are their bottlenecks of production is crucial.
slurry can be cooked at 165°C for 3–5 minutes or at 90–105°C for 1–3 hours
[14, 27], resulting in a mixture with gelatinized starch that is converted to
dextrins and oligosaccharides by -amylase [20, 30].
The output from liquefaction is called corn mash which is rich in
short-chain saccharides, showing partial solubility with no fermentative
characteristic [30]. The corn mash is cooled approximately to 30°C [14,
21], and it proceeds to the saccharification process, in which the conversion
Author Copy
of dextrins into fermentable sugars occurs. This hydrolysis happens in the
presence of glucoamylase (GA, also called amyloglucosidase or AMG) for
converting dextrins into glucose [10, 201]. In the past, the saccharification
was carried by acid hydrolysis in a hazardous operation that occurred in
extreme conditions of temperature and pH with sugar production yields of
85% [20]. The utilization of enzymes led to reduce the disadvantages and
raise the yield up to 95–97% [30].
The corn hydrolysate is subjected to ethanol fermentation in which yeast
Saccharomyces cerevisiae consumes glucose and produces ethanol and
CO2 under anaerobic conditions [1, 31, 200]. The yeast grows in seed tanks
and then is transferred to the sugary syrup in the fermentation tanks [21].
Fermentation is the main operation in a distillery plant. It runs for 42–55 h
using multiple fermenters as batch operation of the plant [32].
There are many different possible configurations for this process. The
most common modes are SHF (separate hydrolysis and fermentation) and
SSF (simultaneous saccharification and fermentation) [202]. The main
benefit of the SSF process is to avoid the osmotic shock to yeast due to
the high glucose concentrations. In this process, glucose is slowly released
through the saccharification process and immediately consumed by the yeast
to produce ethanol, preventing a glucose concentration that can cause inhibi-
tion [33, 203, 204]. The SSF technique can provide up to 8% more ethanol
than the SHF for the same amount of grain [24, 32].
The fermentation product is composed by a liquid portion containing
14–20% of ethanol [29, 199]. Thus, the distillation goal is to obtain a
product with higher ethanol purity (92–98%) [21, 34]. Ethanol is separated
from water and non-fermentable residues which form the stillage [201]. This
process can be separated in two stages. In the first column, unconverted
solids and heavy compounds are removed at the bottom, while the top outlet
feeds the second column with a liquid portion containing 30–40% ethanol. In
the second column, ethanol is purified in which hydrous ethanol is obtained
[32]. Pure ethanol can be recovered by combining distillation and molecular
sieves which capture residual water to achieve anhydrous ethanol [35].
Apple Academic Press
The stillage includes fiber, oil, and protein components of the grain,
as well as non-fermented starch. The stillage can be processed to obtain
byproducts [24]. It can be centrifuged to separate the solid and liquid
fractions and produce a variety of coproducts known as distiller’s grains
[36]. The most popular is distillers dried grains (DDGs) which is used as
animal feed [14]. In addition to animal feed, another coproduct obtained
from the dry grind operational plants is distillers corn oil. It is recovered from
Author Copy
the concentrated stillage liquid portion obtained from the centrifugation of
whole stillage [10, 198].
The requirement of high temperatures is an unfavorable factor in bioethanol
industries, since it increases the energy demand, which adversely affects the
cost-effectiveness of the product, even though the elevated temperatures help
contamination control [37]. In the context of 1G corn ethanol, some of the
research’s address reducing operational costs such as the energy demand [38].
Other issues that also draw attention are crop improvement [5], new enzymes
[3], genetic engineering in yeast strains [39, 40] and coproduct recovery
(Figure 10.1) [41].
In Brazil, this biomass is largely used for 1st generation ethanol [6, 199].
The 1st generation ethanol is produced by processing sugarcane via
fermentation of sucrose-rich juice or molasses [42, 43]. In integrated sugar
and alcohol industries (the most common), there is a strict relation between
ethanol production and other main products (sugar and energy generated
by burning bagasse) which is influenced by market demand, within the
biorefinery concept [44, 45]. Figure 10.2 exemplifies a common industrial
processes of a sugarcane biorefinery (1st generation) with joint production of
ethanol, sugar, and energy.
At the reception, the harvested cane is aimed at cleaning tables and the
feeding system. Furthermore, in this stage the impurities, such as vegetable
and mineral residues are removed. If the harvested cane is chopped, a dry
wash is adopted to minimize the loss of sugars during the reception process.
Then, the clean cane is sent for preparation [46, 195].
stage in a usual process, the chopped cane is crushed via mills (a set of three to
five cylindrical rolls), where the juice is separated from the vegetable fibrous
fraction (bagasse). They are usually sequential mill systems. The last set of
milled sugarcane is soaked into warm water, a way to increase the extraction
of residual sugars [200]. Despite the separation that has been already done,
a fibrous fraction still remains in the juice. In order to remove it, sieves are
used, and the solid material is sent for recirculation in the mill system. The
Author Copy
juice from the first group of mills is often utilized for sugar production, due
to its great degree of purity and concentration of sugars. On the other hand,
the juice of the other groups of mills which are combined with molasses is
used for ethanol production [202–206].
FIGURE 10.1 General scheme for the production of 1G corn ethanol in biorefinery.
Author Copy
FIGURE 10.2 General scheme for the production of 1G ethanol, sugar, and energy from
sugarcane in biorefinery.
As a result, a liquid with the low ethanol content (maximum 10° GL)
is obtained, and CO2 which is still washed in adsorption columns is used
to recover transported ethanol. After fermentation, the juice extracted is
submitted to centrifugation, and the cells are removed in this process,
Apple Academic Press
Author Copy
extraction using monoethylene glycol, or adsorption in sieves [53, 209].
The first generation biorefineries are considered a consolidated tech-
nology with a low-cost production line. However, they are still dependent
on agricultural resources, arable land, climatic conditions, domestic supply
market and investment capital which should be implemented in new regions
or countries [184].
The industrial facilities for the production of 1G ethanol are also designed
for joint production of sugar and energy cogeneration. This has been resulted
from the burning of existing bagasse and straw, as well as the generation of
other products such as biogas obtained from fermented broth and proteins in
the form of yeast cells [13].
The production of sugar within the biorefinery follows common lines for
the production of ethanol. In this approach, the concentration of the treated
juice is adjusted so that heated processes of evaporation of the broth, crystal-
lization, and drying of the sugars occur (humidity between 0.5 and 2%) [52,
205]. Molasses obtained after crystallization can be used to produce ethanol.
In the energy cogeneration system, bagasse (for mechanized harvests) is
burned in a boiler and produced steam is driven to turbines, which in turn
are connected to electric generators [49, 54]. Thus, a part of the generated
electricity can be sold to distributing companies, and the exhausted steam
can also be used in various operations at process units that require thermal
energy [23, 55, 56, 208].
production, affecting the food and land usage concerns [19, 57]. In order to
better apply the feedstock instead of the focus on the increase of its production,
the second-generation biofuel has been a prominent technology since it relies
on lignocellulosic materials [58]. Lignocellulosic substances are variable and
Apple Academic Press
available materials, which can be divided into agricultural residues, forest resi-
dues, municipal solid waste, and energy crops (herbaceous or woody plants)
[2]. The main components of this biomass are cellulose, hemicellulose, and
lignin with small quantities of extractives and ashes [59]. Some examples of
biomass applied for the second-generation bioethanol are sugarcane bagasse
[60], eucalyptus [61], hazelnut shell [62], corncob [63], banana crop [64], rice
straw [65], coconut husk [58], and sugar beet [210]. Each biomass presents
Author Copy
specific quantities of cellulose, hemicellulose, and lignin so that sugarcane
bagasse, for example, is composed of 20–25% lignin, 40–50% cellulose and
30–35% hemicellulose, making it a potential feedstock for 2G biofuel [57,
66]. Another example of potential biomass is sugar beet which is also used
for bioethanol production and presents a high quantity of hemicelluloses
(24–32%), and cellulose (22–30%) with a very low lignin amount [210].
Even though the lignocellulosic residues are abundant and do not interfere
in food production, it is necessary to optimize the biomass bioconversion
in order to produce an economically viable second-generation bioethanol
[66]. Bioconversion of lignocellulosic materials consists of four main steps
including pretreatment, enzymatic hydrolysis, fermentation, and product
recovery [67]. In general, pretreatment contributes to the second-generation
bioethanol production by facilitating the subsequent hydrolysis through the
modification of the amorphous region and porosity of matrix with separating
cellulose from hemicellulose and the lignin [68]. Many technologies have
already been developed for lignocellulose pretreatment, and they can be
divided into physical, chemical, physical-chemical, and biological [211].
Each type of the pretreatment has demonstrated advantages and disadvan-
tages, acting in different methods to facilitate cellulose hydrolysis.
A large number of technologies have been focusing on the development
of an efficient and economic lignocellulosic pretreatment. Pretreatment is
one of the main macro-steps of the process in biorefineries and is closely
related to the subsequent stages of hydrolysis and fermentation. Therefore,
it is necessary to use a method that presents, in addition to high efficiency,
low energy consumption and reduced chemical catalysts to result in a high
recovery of carbohydrate fractions, low or no formation of fermentation
inhibitors and high simplicity for use in a larger scale.
Physical methods include milling [69], extrusion [70], and irradiation
[71]. Different methods of irradiation such as gamma rays and microwave
Author Copy
Other examples of combinations are steam explosion (SE) [75], ammonia
recycle percolation [214], microwave-chemical [76], and hydrodynamic
cavitation (HC) [77]. For example, in an optimized condition of NaOH
pretreatment, a maximum glucose production equivalent to 85% of enzymatic
digestibility is observed. For the pretreatment of CH, it is possible that this
number reaches 96% digestibility. This is a new solution for the bottleneck
regarding 2 G produce ethanol [78]. Pretreatment based on HC has advantages
compared to other methods, such as the requirement of milder pretreatment
conditions and shorter process times [77]. Microorganisms such as fungi
and bacteria are also used in lignocellulosic pretreatment. They are capable
of modifying and degrading the complex structure into simpler substrates
by enzyme digestion. Some examples of microorganism are white-rot,
brown-rot, and soft-rot fungi. White-rot fungi stand out for providing
better sugar yields [215, 216]. However, microbiological pretreatment has
considerable disadvantages, especially the low hydrolysis rate caused by the
presence of inhibitors and the broth conditions [79].
After the pretreatment, it is necessary to convert the polymeric carbo-
hydrate obtained from lignocellulosic material into fermentable sugars
(Figure 10.3), due to the yeast inability to process carbohydrate polymers
[217]. Therefore, two hydrolysis methods can be applied in order to provide
second-generation bioethanol production, including acid and enzymatic
hydrolysis, which are chosen according to the biomass composition [67].
Acid hydrolysis can be performed by dilute acid or concentrated acid with
demanding different temperature and pressure conditions [17]. Sulfuric acid
(H2SO4) [80] is mostly applied for acid hydrolysis, however, researchers have
also developed different methods, including hydrochloric acid (HCl), nitric
acid (HNO3), and phosphoric acid (H3PO4) [81]. Enzymes can also be used to
hydrolyze both cellulose and hemicellulose components, in order to produce
fermentable sugars. Enzymatic hydrolysis (enzymatic saccharification) has
Author Copy
286 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
arabinose) and hexoses (e.g., glucose and mannose) [67, 218]. A co-culture
system has been reported in the literature, favoring both pentose and hexoses
fermentation. Farias and Filho [84] evaluated a co-culture fermentation using
hexoses-fermenting yeasts (such as Saccharomyces cerevisiae) and xylose-
fermenting yeasts (Scheffersomyces stipitis and Spathaspora passalidarum),
resulting in a maximum ethanol titer of 49.2 gL−1. Four main alternative
processes are commonly developed for bioethanol production, namely SSF,
SHF, SSCF, and consolidated bioprocessing (CBP) [7]. Each alternative
interferes in the process with advantages and disadvantages, giving the
lignocellulosic bioethanol production options according to a variety of the
processes, especially by consideration of the feedstock, microorganism, and
final product [41, 208]. Due to the many advantages of SSCF process, this
process was investigated for sugarcane bagasse pretreated with alkali assisted
HC. With the developed methodology, 62.33% of the total hydrolyzed carbo-
hydrate and 17.26 g/L of ethanol production were obtained, indicating 0.48
g of ethanol/g of glucose and xylose consumed by Scheffersomyces stipitis
NRRL-Y7124 [219].
Overall, the use of LCB for the second-generation bioethanol has
presented many advantages. The sustainable pathway, fossil fuel substitu-
tion, and the use of a variety of feedstock are some benefits of the second-
generation ethanol production [19]. Despite technological barriers, plants
of 2G bioethanol are already operating in the world [6]. Some examples
of the second-generation commercial ethanol plants are GranBio, Raízen,
Poet-DSM, BetaRenewables, Abengoa (plant bought by Synata Bio), and
DuPont [85]. Among them, two Brazilian companies, namely GranBio and
Raizen have reached commercial scale by predominantly using sugarcane
bagasse as the LCB [86]. In Nevada, Iowa, USA, the cellulosic ethanol plant
of DuPont has developed a projected production of 30 million US gallons of
bioethanol from corn stover as a lignocellulosic material [87].
Author Copy
biorefinery was built in March 2017. In this first plant, the ethanol produced
up to 5 million gallons of cellulosic ethanol per year was attained using this
technology [185]. In the corn ethanol refineries, this form of ethanol produc-
tion is combined with the 1G. Moreover, the 1.5G technology is still attractive
for the usage of cellulosic raw material, and it allows a reduction in GHG
emissions and indulges companies with producer titles. Therefore, renewable
products can receive technological and government incentives [14, 90].
The corn grain fiber, which is used in biorefineries for the production of
1.5 G ethanol, has some particular characteristics. Firstly, it includes its pure
and homogeneous composition formed by cellulose, hemicellulose, residual
starch, and little lignin, being a minimally recalcitrant source. Furthermore,
this material requires fewer pretreatment steps which support the diminishing
of the coast of ethanol production [220, 222].
Production of 1.5 G integrates a process that transforms corn fiber to
cellulosic ethanol with existing ethanol plants. In this way, the pathway to
cellulosic ethanol is accomplished by combining mechanical, chemical,
and biological processes. The fiber stream is submitted to an acid pretreat-
ment that deconstructs it; thus, it can access the cellulose. Additionally,
the cellulose stream is broken down into sugars with the enzyme cocktail
(Novozymes), and then the C5 and C6 sugars are transformed into ethanol
with other process detailed in the following parts [8].
The 1.5 G process was developed through the collaborations with the two
world-leading biotechnology companies. Nowadays, two different ways are
adopted for the bioconversion of the corn grain fiber into ethanol. One way is
to adopt the process separately, where the hydrolysis of starch and grain fiber
occurs separately. On the other hand, a second way is to adopt combined
systems of pretreatment of the grain (starch and fiber) before fermenta-
tion [14]. The companies that start this process have with ICM’s patented
technologies, selective milling technology (SMT), and fiber separation
technology (FST). SMT selectively grinds corn slurry to make the starch and
oil more accessible in the entire process. Novozymes provided the enzyme
cocktail which converts the cellulose stream into accessible sugars. DSM has
developed yeast that ferment both the C5 and C6 sugars.
Apple Academic Press
Author Copy
presented in the aqueous fraction after distillation/dehydration, which are
still utilized for animal feed [222].
The 1.5 G technology demonstrates a great potential for ethanol generation,
and its production process has a great potential to be adapted in a smoothie
way for the facilities already existing in a corn starch ethanol plant [14]. Some
future considerations can be suggested for the improvement of lignocellu-
losic bioethanol, such as process optimization to minimize water and energy
usage, the development of engineered microorganisms and the combination
of different microorganisms, which increases the fermentation capacity for
bioethanol production [91]. In addition to these strategies, a significant reduc-
tion in the enzymes and pretreatment costs is necessary to improve the yield
and productivity of the bioethanol for increasing its commercial scales around
the world [92, 93]. Therefore, integration, and consolidation to improve the
whole second-generation bioethanol production will make this a competitive
market with petroleum fuel in the biorefinery process.
Even with the high performance of the ethanol production process, there is still
a way for top countries such as Brazil to increase its productivity even more.
This could occur due to the approach of several crops with higher productivity
in terms of ethanol per unit of area [94]. For example, in Brazil, sugarcane
can be produced between 6,000–7,500 L/ha, compared to sugar beet (EU)
5,500 L/ha and corn (USA) 3,800 L/ha [95–97]. This higher productivity of
sugarcane implies the lower ethanol production costs. However, even with
a harvest cycle higher than that of corn or sugar beet, there is an off-season,
in which sugarcane cannot be processed, since it is not possible to store it.
Nevertheless, this is different from corn, which can be stored all year round
[96]. Both biomass has differences in relation to their derived products. For
example, in addition to ethanol, sugarcane gives a rise in electricity and
sugar. This underproduction has a strategic value for the sugarcane mills, as it
Apple Academic Press
allows the capture of value in different markets [48]. In this context, corn can
raise ethanol and other food products, such as oil and protein for animal feed
(BBGDS). According to RFA [225], for each corn unit converted to ethanol,
a third part returns to the animal nutrition. However, corn bioprocess plant is
not self-sufficient in energy terms as the Brazilian one.
Therefore, con is the biggest gain for the Brazilian bioenergy process to
invest in new biorefinery models, such as the "flex" ones, which are capable
Author Copy
of integrating two different feedstocks, namely corn and sugarcane, or the
use of their residues as an alternative feedstock [9]. In the last case, one
strategy for biorefinery model occurs when it integrates residues of second-
generation biomass in first-generation process, such as sugarcane bagasse
and straw, which could help the sugar-energy sector to overcome low levels
of profitability of ethanol [98].
Fuel ethanol production generally consists of five-stage process: (i) raw
material collection; (ii) pretreatment; (iii) hydrolysis; (iv) fermentation; (v)
separation and (vi) dehydration (downstream). However, there are some
differences when it is produced from different feedstocks [99]. In the case of
flex biorefineries, some steps are adapted and currently these are key points
of challenges to increase conversion yields. Some of these integrations and
flex modes are presented in Figure 10.4.
Moreover, the application of alternative biorefinery processes help to
reduce the momentum for investments in the construction of new plants in
the region. At the same time, this situation needs intense search for innova-
tions with the potential to increase the profitability levels of these young
companies. This incentive could increase the production of corn ethanol
in flex mills to 1.3 billion liters [226]. Other phenomenon that this process
integrates is fermentation of sugarcane juice (or molasses) and starchy
feedstocks, providing a number of advantages over conventional distilleries
[94]. This process has a faster fermentation (34–36 h) in comparison with
the traditional process (45–60 h) adopted by distilleries in the USA. In this
approach, less sugar is deviated for yeast multiplication and production of
cellular biomass [227]. Thus, this mode of “flex” biorefinery uses the same
distillation system applied for sugarcane and extends the period of ethanol
production to 345 days a year, reducing initial investments and fixed costs.
For this new process each corn ton allows to produce 415 L of ethanol and
250 kg of DDGS (dry distillation grain) [94, 96, 228].
Author Copy
Integrated Production of Ethanol from Starch and Sucrose 291
between them [9]. Flex-fuel plants of corn and sugarcane have an energy
balance and reduction of GHG emissions that do not affect the fermenta-
tion performance. This process is economically viable in regions with corn
supply at low prices and high demand of DDGS for animal feed. Thus, it is
an opportunity for Brazilian biorefineries in corn producing regions [96].
In addition, ethanol production from starchy gives two feedstock options to
reduce the end problem of biofuels, and to enable the use of LCB-derived
Author Copy
sugars for the other best production [23, 100, 227].
In the case of lignocellulosic materials, the cellulose and hemicellulose
content account for more than 60% of dry weight (ex. sugarcane bagasse)
and can be converted into fermentable sugars either by acid or enzymatic
hydrolysis [7, 100]. These sugars could provide substrate to produce not only
ethanol, but also other biofuels such as butanol and 1-methlypropanol. All
those strategies could make the possible development of a new economically
and energetically panorama for flex biorefineries [227].
For example, in Brazil, there are two industrial plants integrated
into operation for the production of second-generation ethanol that use a
semi-industrial process. These processes were originated from investment
programs of the Brazilian government in the second-generation ethanol
launched in 2011 [96, 101]. Nevertheless, the production costs of the second-
generation ethanol are still high, mainly concerning equipment handling at
bagasse pretreatment and the use of enzymes [96, 102].
Comparing characteristics between the different types of biomass,
researchers have found that sugarcane is currently the most promising
source for the production of biofuels [229]. However, the flex biorefinery
shows that it is possible to have one process more promising that sugarcane
biorefinery [94, 96]. Table 10.2 shows different situations corresponding to
the first- and second-generation plant scenarios dedicated exclusively to 1G
processing (situation 1), and corn processing (situation 2). This case shows
consolidated ethanol production in the world with different biomass. Situa-
tion 3 shows the lowest amounts of GHG compared to all other processes
studied in Table 10.2. Moreover, this situation could impact in a positive
way, when corn is integrated into a sugarcane flex plant. The main factor that
contributes to decrease GHG emissions is the inclusion of corn in integrated
biorefineries (situation 3 and 4) which in this process exists the higher ethanol
Author Copy
TABLE 10.2
Apple Academic Press
(Continued)
Biorefinery Process Characteristics, Advantages, and Limitations References
5. Corn and sugarcane ethanol • In this biorefinery, corn is stored and used together with sugarcane to produce ethanol for 8 [48, 96, 108]
integrated production months per year.
off-season (Flex-Example 3) • This type of biorefinery, with the integration of the sugarcane and corn culture, can
increase the production of Brazilian ethanol.
6. Flex sugarcane 1 G, 2 G • Cellulosic materials such as bagasse which are already present in the 1G ethanol production [7, 48, 103,
chain could be used as raw material for 2G production by conventional infrastructure. 105]
• High pretreatment price
• High enzyme cost
• Low pentose fermentation yield
Integrated Production of Ethanol from Starch and Sucrose
Author Copy
294 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
Furthermore, the situation 6 could reduce the investments in the construc-
tion of new plants in the world, because of the use of 2 G biomass [103]. At
the same time, this situation sparks an intense research for innovations with
a potential to increase the profitability levels of these young sectors, and for
the investment in more flex bioprocess [104]. One disadvantage is that this
process needs an overwork on the logistics side, because freight and storage
is not a simple issue for industry [13]. In addition, the 1G + 2G configuration
has the greatest efficiency (44.4%) compared to 1G biorefinery configuration
[105]. Another problem for this situation is that investors demand a return of
5-fold investment. However, the most important factor in situation 6 is the low
investment (50 million to adapt an existing biorefinery) of these biorefineries,
which opens many possibilities for developing countries [14, 103].
Currently, the demand for biofuels has increased due to environmental, social,
and technological care [13]. In this view, there is a great worldwide interest
in the use of biomass residues, such as cellulosic residues (for example,
sugarcane bagasse and corn straw) for the production of biofuels 2 G [7, 103].
In the context of scaling up production, there is the integrated biorefinery
configuration, called flex biorefinery, in which there is a possibility of integrating
different types of biomass for the production of 1G ethanol. For example, in
Brazil, there is a high rate of productivity in this biorefinery configuration during
a four-month period in which there is no ethanol production, due to the lack of
raw material and thus the plant goes through a long period of maintenance [104,
111]. This fact occurs because of the perishability of sugarcane [232]. As seen
before, this problem is not common for corn, and the best option is the use of
both feedstocks for ethanol production [23, 105].
All biorefineries have the potential to meet the demand for the produc-
tion of bioethanol, but there are still many challenges in the integration of
Apple Academic Press
Author Copy
little protein or other nutrients (corn) that fermentation microbes require,
however, not all feedstocks have these advantages, and it is important to add
some supplementation [98, 103].
Other challenges that need to be considered are harvest, transportation
costs and storage. For example, sugar crops (sugarcane) are especially difficult
to store without losses, and this might also be a problem with lignocellulosic
feedstocks [104]. Another challenge in the process is the stage of saccharifica-
tion, which is important because current enzymes for hydrolysis are more
expensive and less effective than corresponding enzymes used with starch
[111, 208]. Another alternative for saccharification process is to conduct SSF
of lignocellulose due to the feedback inhibition of the enzymes [23, 197].
Pretreatment is the most significant bottleneck in biorefinery process;
however, pretreatment is not needed for all sugar crops. It is used for
lignocellulosic materials where it is an overriding feedstock. However, the
lignocellulose pretreatments are often sufficiently intensive and generate
inhibitory compounds, and this is prejudicial to the total process. In addition,
it is difficult for the industrial process, since pretreatments have high costs
and do not have short process time options to be applied for the continuous
industrial process [98, 112, 113].
Another challenge is that microbial growth is limited due to the presence
of inhibitors in the fermentation process generated in previous stages
(pretreatment). In addition, different sugars (C5 and C6 sugars) in substrates
obtained from LCB are difficult to be metabolized by many wild type
microorganisms because it comes from pretreated biomass. In fermentation,
there are difficulties with the concentration of sugars and solids content in
the medium which are also unsatisfactory factors for the system [111, 236].
The cost of integrated 1G + 2G biorefineries requires two to four times
more investment than 2G ethanol prices to ensure competitiveness, which is
one of the main problems in flex production. However, this can be fixed when
to its cost is the heat exchange process which can be reduced by adopting
developed schemes for evaporation and backward integrated evaporation
that is a new perspective for flex biorefinery [98, 241–244].
1G and 2G biorefineries have higher investments in costs than 1G+ 2G
+ electricity integration, due to the hydrolysis step with cogeneration which
presents a lower yield. The 1G + electricity schemes have provided better
results and it is superior to 1G+2G + electricity equivalents, mainly because
Author Copy
of its highest investment costs [234, 241]. Other problems of sugarcane 1 G
and 2 G plants are related to enzyme parameter costs and reduced electricity
production. However, the advantage of this process lies in ethanol prices with
lower capital costs, presenting the greatest potential to increase the economic
utilization of 1G + 2G (flex) schemes. In this regard, the cost is lower than
2G biorefinery process and it has greater productivity than 1G [197].
Other option in the biorefinery integration is the development of plants with
corn and sugarcane as feedstocks that results in 1G and 2G processes. This
biorefinery could have high productivity which could help solving the global
demand of ethanol [112, 245]. Thus, sugarcane biomass is more versatile to be
used in this kind of biorefinery, because it could be integrated with different
biomasses (sugarcane and corn) in 1G or 2G integrated [98, 241, 246].
Energy integration in a sugarcane biorefinery can provide economical
advantage, environmental benefits, and increased ethanol production. The
last factor is related to the lower steam consumption in the plant due to
energy integration and consequently, and less bagasse needs to be burnt in
the electricity generation. Hence, its surplus can be made available for the
production of second-generation ethanol [51, 115, 234].
There are other integration processes with different advantages, such as,
1G, 2G and cogeneration system (1G+2G+COGEN), 1G and 2G from corn,
1G and 2G for sugarcane, 1G (biomass A) and 2G (biomass B) that could be
the future of fuels from biomass [111, 113]. These integrations could give
us positive results with 2G, which involves the production of 1G and 2G
ethanol in a combined distillery, enzymatic hydrolysis, and cogeneration
plant [112]. These kind of biorefineries offer many techno-economic and
environmental alternatives.
Total production cost calculations for these kinds of biorefineries can be
resulted in 74 settings, covering 5 fuel output types, 8 feedstock types, 12
10.7 CONCLUSION
Author Copy
First and second-generation bioethanol production from sugarcane and corn
are successful processes established in several countries, and it is currently
analyzed to find out the possibility for the integration of other biorefinery
modes to enhance some particular bottlenecks. Those bottlenecks are the
seasonal lack of the feedstocks and approach of second-generation sugars.
Several strategies, such as the alternation of feedstocks in off-seasons and
the integration of second-generation into the first-generation processes are
deeply evaluated. However, the integration of these modes in biorefinery is a
hard-analytical work to understand the role that plays each step-in bioethanol
production from two feedstocks with differences in structure, sugar content,
and processing. Even though the correct application of those strategies
could result in a successful integration from both biomass, the main steps in
bioprocesses such as pretreatments, saccharification, and fermentation need
more work to obtain a cost-effective competitive conversion.
KEYWORDS
• consolidated bioprocessing
• distillers dried grains
• hydrochloric acid
• saccharification and fermentation
• separate hydrolysis and fermentation
• simultaneous saccharification and co-fermentation
REFERENCES
1. Bai, F. W., Anderson, W. A., & Moo-Young, M., (2008). Ethanol fermentation
technologies from sugar and starch feedstocks. Biotechnology Advances, 26(1), 89–105.
Apple Academic Press
2. Zabed, H., Sahu, J. N., Suely, A., Boyce, A. N., & Faruq, G., (2017). Bioethanol
production from renewable sources: Current perspectives and technological progress.
Renewable and Sustainable Energy Reviews, 71, 475–501.
3. Akram, F., Ul Haq, I., Imran, W., & Mukhtar, H., (2018). Insight perspectives of
thermostable endoglucanases for bioethanol production: A review. Renewable Energy,
122, 225–238.
4. Mussatto, S. I., Machado, E. M., Martins, S., & Teixeira, J. A., (2011). Production,
composition, and application of coffee and its industrial residues. Food and Bioprocess
Author Copy
Technology, 4(5), 661.
5. Gumienna, M., Szwengiel, A., Lasik, M., Szambelan, K., Majchrzycki, D., Adamczyk,
J., & Czarnecki, Z., (2016). Effect of corn grain variety on the bioethanol production
efficiency. Fuel, 164, 386–392.
6. Vallejos, M. E., Kruyeniski, J., & Area, M. C., (2017). Second-generation bioethanol
from industrial wood waste of South American species. Biofuel Research Journal, 4(3),
654–667.
7. Milanez, A. Y., Nyko, D., Valente, M. S., Xavier, C. E. O., Kulay, L. A., Donke, A.
C. G., & Capitani, D. H. D., (2014). Ethanol production by integrating off-season
corn into mills Final report 2019/20 crop, south-central region of Brazil. Sugarcane:
Environmental, Economic Assessment and Policy Suggestions. https://web.bndes.gov.
br/bib/jspui/handle/1408/1921 (accessed on 28 October 2021).
8. Sydney, E. B., Letti, L. A. J., Karp, S. G., Sydney, A. C. N., De Souza, V. L. P., De
Carvalho, J. C., & Soccol, C. R., (2019). Current analysis and future perspective of
reduction in worldwide greenhouse gases emissions by using first and second-generation
bioethanol in the transportation sector. Bioresource Technology Reports, 7, 100234.
9. Cabral, F. F. R., (2019). Agronomic performance and nutritional balance in the corn
plant fertigated with concentrated vinasse and potassium chloride.
10. Kumar, D., & Singh, V., (2019). Bioethanol production from corn. In: Corn (pp. 615–631).
AACC International Press.
11. CONAB, (2019). Retrieved from https://www.conab.gov.br// (accessed on 28 October
2021).
12. Vanier, N. L., Vamadevan, V., Bruni, G. P., Ferreira, C. D., Pinto, V. Z., Seetharaman, K.,
& Berrios, J. D. J., (2016). Extrusion of rice, bean and corn starches: Extrudate structure
and molecular changes in amylose and amylopectin. Journal of Food Science, 81(12),
E2932–E2938.
13. Guo, M., Song, W., & Buhain, J., (2015). Bioenergy and biofuels: History, status, and
perspective. Renewable and Sustainable Energy Reviews, 42, 712–725.
14. Mohanty, S. K., & Swain, M. R., (2019). Bioethanol production from corn and wheat: Food,
fuel, and future. In: Bioethanol Production from Food Crops (pp. 45–59). Academic Press.
15. EMBRAPA, (2019). Retrieved from: https://www.embrapa.br/agua-na-agricultura/links
(accessed on 8 December 2021).
16. Zhao, L., Ou, X., & Chang, S., (2016). Life-cycle greenhouse gas emission and energy
use of bioethanol produced from corn stover in China: Current perspectives and future
prospectives. Energy, 115, 303–313.
17. Torabi, S., Satari, B., & Hassan-Beygi, S. R., (2020). Process optimization for dilute
acid and enzymatic hydrolysis of waste wheat bread and its effect on aflatoxin fate and
ethanol production. Biomass Conversion and Biorefinery, 1–9.
18. Agarwal, M., Rampure, M., Todkar, A., & Sharma, P., (2019). Ethanol from maize: An
Apple Academic Press
Author Copy
D., Cantarella, H., & Franco, H. C. J., (2017). Agronomic and environmental implications
of sugarcane straw removal: A major review. GCB Bioenergy, 9(7), 1181–1195.
23. Sharma, B., Larroche, C., & Dussap, C. G., (2020). Comprehensive assessment of 2G
bioethanol production. Bioresource Technology, 123630.
24. Bothast, R. J., & Schlicher, M. A., (2005). Biotechnological processes for conversion of
corn into ethanol. Applied Microbiology and Biotechnology, 67(1), 19–25.
25. Eskicioglu, C., Kennedy, K. J., Marin, J., & Strehler, B., (2011). Anaerobic digestion
of whole stillage from dry-grind corn ethanol plant under mesophilic and thermophilic
conditions. Bioresource Technology, 102(2), 1079–1086.
26. Yu, J., Xu, Z., Liu, L., Chen, S., Wang, S., & Jin, M., (2019). Process integration
for ethanol production from corn and corn stover as mixed substrates. Bioresource
Technology, 279, 10–16.
27. Lamsal, B. P., Wang, H., & Johnson, L. A., (2011). Effect of corn preparation methods
on dry-grind ethanol production by granular starch hydrolysis and partitioning of spent
beer solids. Bioresource Technology, 102(12), 6680–6686.
28. Wang, M., Wu, M., & Huo, H., (2007). Life-cycle energy and greenhouse gas emission
impacts of different corn ethanol plant types. Environmental Research Letters, 2(2), 024001.
29. Zhao, Y., Damgaard, A., & Christensen, T. H., (2018). Bioethanol from corn stover-a
review and technical assessment of alternative biotechnologies. Progress in Energy and
Combustion Science, 67, 275–291.
30. Power, R. F., (2003). Enzymatic conversion of starch to fermentable sugars. The Alcohol
Textbook, 23–32.
31. Bonassa, G., Schneider, L. T., Cremonez, P. A., De Oliveira, C. D. J., Teleken, J. G., &
Frigo, E. P., (2015). < b> Optimization of first-generation alcoholic fermentation process
with< i> Saccharomyces cerevisiae. Acta Scientiarum. Technology, 37(3), 313–320.
32. Saville, B. A., Griffin, W. M., & MacLean, H. L., (2016). Ethanol Production technologies
in the Us: Status and future developments. In: Global Bioethanol (pp. 163–180). Academic
Press.
33. Chen, H., & Li, G., (2013). An industrial level system with nonisothermal simultaneous
solid state saccharification, fermentation and separation for ethanol production.
Biochemical Engineering Journal, 74, 121–126.
34. Madson, P. W., (2003). Ethanol distillation: The fundamentals. The Alcohol Textbook, 4.
35. Chum, H. L., Warner, E., Seabra, J. E., & Macedo, I. C., (2014). A comparison of
commercial ethanol production systems from Brazilian sugarcane and US corn. Biofuels,
Bioproducts and Biorefining, 8(2), 205–223.
36. Sharma, A., & Bhargava, R., (2016). Production of Biofuel (Ethanol) from Corn and
co-product evolution: A review. IRJET, 3(12), 745–749.
37. Cripwell, R. A., Favaro, L., Viljoen-Bloom, M., & Van, Z. W. H., (2020). Consolidated
bioprocessing of raw starch to ethanol by Saccharomyces cerevisiae: Achievements and
Apple Academic Press
Author Copy
41. Sharma, V., Sharma, S., & Kuila, A., (2016). A review on current technological
advancement of lignocellulosic bioethanol production. Journal of Applied Biotechnology
Bioengineering, 1(2), 61–66.
42. Klein, B. C., Chagas, M. F., Watanabe, M. D. B., Bonomi, A., & Maciel, F. R., (2019).
Low carbon biofuels and the new Brazilian national biofuel policy (RenovaBio): A case
study for sugarcane mills and integrated sugarcane-microalgae biorefineries. Renewable
and Sustainable Energy Reviews, 115, 109365.
43. Emori, E. Y., Ferreira, J., Secchi, A. R., Ravagnani, M. A., & Costa, C. B., (2020).
Dynamic study of the evaporation stage of an integrated first and second-generation
ethanol sugarcane biorefinery using EMSO software. Chemical Engineering Research
and Design, 153, 613–625.
44. Moraes, M. A. F. D., Oliveira, F. C. R., & Diaz-Chavez, R. A., (2015). Socio-economic
impacts of Brazilian sugarcane industry. Environmental Development, 16, 31–43.
45. Júnior, J. C. F., Palacio, J. C. E., Leme, R. C., Lora, E. E. S., Da Costa, J. E. L., Reyes,
A. M. M., & Del, O. O. A., (2020). Biorefineries productive alternatives optimization in
the Brazilian sugar and alcohol industry. Applied Energy, 259, 113092.
46. Junqueira, T. L., Dias, M. O. S., Jesus, C. D. F., Mantelatto, P. E., Cunha, M. P., Cavalett,
F. O., & Bonomi, A., (2011). Simulation and evaluation of autonomous and annexed
sugarcane distilleries. Chemical Engineering Transactions, 25, 941–946.
47. Dias, F. M. D. S., RM, M. P., Cavalett, O., Rossell, C. E. V., Bonomi, A., & Leal, M.
R. L. V., (2015). Sugarcane processing for ethanol and sugar in Brazil. Environmental
Development, 15(2015), 35–51.
48. Cavalett, O., Junqueira, T. L., Dias, M. O., Jesus, C. D., Mantelatto, P. E., Cunha, M. P., &
Bonomi, A., (2012). Environmental and economic assessment of sugarcane first generation
biorefineries in Brazil. Clean Technologies and Environmental Policy, 14(3), 399–410.
49. Khatiwada, D., Leduc, S., Silveira, S., & McCallum, I., (2016). Optimizing ethanol and
bioelectricity production in sugarcane biorefineries in Brazil. Renewable Energy, 85,
371–386.
50. Andrade, L. P., Crespim, E., De Oliveira, N., De Campos, R. C., Teodoro, J. C., Galvão,
C. M. A., & Maciel, F. R., (2017). Influence of sugarcane bagasse variability on sugar
recovery for cellulosic ethanol production. Bioresource Technology, 241, 75–81.
51. Macrelli, S., Mogensen, J., & Zacchi, G., (2012). Techno-economic evaluation of 2nd
generation bioethanol production from sugarcane bagasse and leaves integrated with the
sugar-based ethanol process. Biotechnology for Biofuels, 5(1), 22.
52. Chandel, A. K., Junqueira, T. L., Morais, E. R., Gouveia, V. L. R., Cavalett, O., Rivera,
E. C., & Da Silva, S. S., (2014). Techno-economic analysis of second-generation
ethanol in Brazil: Competitive, complementary aspects with first-generation ethanol.
In: Biofuels in Brazil (pp. 1–29).
Apple Academic Press
53. Junqueira, T. L., Dias, M. O., Maciel, F. R., Maciel, M. R., & Rossell, C. E., (2009).
Simulation of the azeotropic distillation for anhydrous bioethanol production: Study on
the formation of a second liquid phase. In: Computer-Aided Chemical Engineering (Vol.
27, pp. 1143–1148). Elsevier.
54. Khatiwada, D., Seabra, J., Silveira, S., & Walter, A., (2012). Power generation from
sugarcane biomass—A complementary option to hydroelectricity in Nepal and Brazil.
Energy, 48(1), 241–254.
55. Serra, L. M., Lozano, M. A., Ramos, J., Ensinas, A. V., & Nebra, S. A., (2009).
Polygeneration and efficient use of natural resources. Energy, 34(5), 575–586.
Author Copy
56. Leal, M. R. L., Walter, A. S., & Seabra, J. E., (2013). Sugarcane as an energy source.
Biomass Conversion and Biorefinery, 3(1), 17–26.
57. Bezerra, T. L., & Ragauskas, A. J., (2016). A review of sugarcane bagasse for second-
generation bioethanol and biopower production. Bioproducts and Biorefining, 10(5),
634–647.
58. Bolivar-Telleria, M., Turbay, C., Favarato, L., Carneiro, T., De Biasi, R. S., Fernandes,
A. A. R., & Fernandes, P., (2018). Second-generation bioethanol from coconut husk.
BioMed Research International.
59. Siqueira, J. G. W., Rodrigues, C., De Souza, V. L. P., Woiciechowski, A. L., & Soccol, C.
R., (2020). Current advances in on-site cellulase production and application on lignocel-
lulosic biomass conversion to biofuels: A review. Biomass and Bioenergy, 132, 105419.
60. González-Bautista, E., Alarcón-Gutiérrez, E., Dupuy, N., Gaime-Perraud, I., Ziarelli,
F., Foli, L., & Farnet-Da-Silva, A. M., (2020). Preparation of a sugarcane bagasse-
based substrate for second-generation ethanol: Effect of pasteurization conditions on
dephenolization. Renewable Energy.
61. Schneider, W. D. H., Fontana, R. C., Baudel, H. M., De Siqueira, F. G., Rencoret,
J., Gutiérrez, A., & Dillon, A. J. P., (2020). Lignin degradation and detoxification
of eucalyptus wastes by on-site manufacturing fungal enzymes to enhance second-
generation ethanol yield. Applied Energy, 262, 114493.
62. Uyan, M., Alptekin, F. M., Cebi, D., & Celiktas, M. S., (2020). Bioconversion of hazelnut
shell using near critical water pretreatment for second-generation biofuel production.
Fuel, 273, 117641.
63. Kleingesinds, E. K., José, Á. H., Brumano, L. P., Silva-Fernandes, T., Rodrigues, Jr. D.,
& Rodrigues, R. C., (2018). Intensification of bioethanol production by using Tween
80 to enhance dilute acid pretreatment and enzymatic saccharification of corncob.
Industrial Crops and Products, 124, 166–176.
64. Guerrero, A. B., Ballesteros, I., & Ballesteros, M., (2018). The potential of agricultural
banana waste for bioethanol production. Fuel, 213(1), 176–185.
65. Fonseca, B. G., Mateo, S., Moya, A. J., & Roberto, I. C., (2018). Biotreatment
optimization of rice straw hydrolyzates for ethanolic fermentation with Scheffersomyces
stipitis. Biomass and Bioenergy, 112, 19–28.
66. Anwar, Z., Gulfraz, M., & Irshad, M., (2014). Agro-industrial lignocellulosic biomass a
key to unlock the future bio-energy: A brief review. Journal of Radiation Research and
Applied Sciences, 7(2), 163–173.
67. Fennouche, I., Khellaf, N., Djelal, H., & Amrane, A., (2019). An effective acid pretreat-
ment of agricultural biomass residues for the production of second-generation bioethanol.
SN Applied Sciences, 1(11), 1460.
68. Aditiya, H. B., Mahlia, T. M. I., Chong, W. T., Nur, H., & Sebayang, A. H., (2016).
Apple Academic Press
Author Copy
20 sorbitan monolaurate. Applied Biochemistry and Biotechnology, 167, 377–393.
72. Kristiani, A., Effendi, N., Styarini, D., Aulia, F., & Sudiyani, Y., (2016). The effect of
pretreatment by using electron beam irradiation on oil palm empty fruit bunch. Atom
Indonesia, 42(1), 9–12.
73. Marinho, C. C. P., Pereira, L. V. B., & Vasconcelos, S. M., (2020). Acid pretreatment of
sugarcane straw aiming at the second-generation ethanol production. Revista Tecnologia
e Sociedade. 16(41), 1–14.
74. Chopda, R., Ferreira, J. A., & Taherzadeh, M. J., (2020). Biorefining oat husks into
high-quality lignin and enzymatically digestible cellulose with acid-catalyzed ethanol
organosolv pretreatment. Processes, 8(4), 435.
75. Equihua-Sanchez, M., & Barahona-Perez, L. F., (2019). Physical and chemical
characterization of Agave tequilana bagasse pretreated with the ionic liquid 1-ethyl-3-
methylimidazolium acetate. Waste and Biomass Valorization. 10, 1285–1294.
76. Damay, J., Duret, X., Ghislain, T., Lalonde, O., & Lavoie, J. M., (2018). Steam explosion
of sweet sorghum stems: Optimization of the production of sugars by response surface
methodology combined with the severity factor. Industrial Crops and Products, 111,
482–493.
77. Mikulski, D., & Kłosowski, G., (2020). Microwave-assisted dilute acid pretreatment in
bioethanol production from wheat and rye stillages. Biomass and Bioenergy, 136, 105528.
78. Terán-Hilares, R., Dionízio, R. M., Prado, C. A., Ahmed, M. A., Da Silva, S. S., & Santos, J.
C., (2019). Pretreatment of sugarcane bagasse using hydrodynamic cavitation technology:
Semi-continuous and continuous process. Bioresource Technology, 290, 121777.
79. Terán-Hilares, R., Dionízio, R. M., Muñoz, S. S., Prado, C. A., De Sousa, J. R., Da
Silva, S. S., & Santos, J. C., (2020). Hydrodynamic cavitation-assisted continuous
pre-treatment of sugarcane bagasse for ethanol production: Effects of geometric
parameters of the cavitation device. Ultrasonics Sonochemistry, 63, 104931.
80. Kucharska, K., Rybarczyk, P., Hołowacz, I., Łukajtis, R., Glinka, M., & Kamiński,
M., (2018). Pretreatment of lignocellulosic materials as substrates for fermentation
processes. Molecules, 23(11), 2937.
81. Gebrehiwot, G., Birhane, K., & Gebrekidan, T., (2020). Optimization of sulphuric acid
hydrolysis process for fermentable sugars from lignocellulosic content of wood sawdust
for production of cellulosic ethanol. Emerging Trends in Chemical Engineering, 7(1).
82. Horst, D. J., Behainne, J. J. R., & Petter, R. R., (2011). Analysis of hydrolysis yields by
using different acids for bioethanol production from Brazilian woods. XVII International
Conference on Industrial Engineering and Operations Management. 1–14.
83. Anindyawati, T., Triwahyuni, E., Maryana, R., & Sudiyani, Y., (2020). The enzymatic
process of lignocellulosic biomass for second generation bioethanol production, the benefits
and challenges: a review. International Journal of Agricultural Technology, 16(3), 529–544.
84. Kim, J. K., Yang, J., Park, S. Y., Yu, J. H., & Kim, K. H., (2019). Cellulase recycling in
Apple Academic Press
Author Copy
vectors to boost the development of biofuels. Industrial Crops and Products, 129, 201–205.
88. Lynd, L. R., Liang, X., Biddy, M. J., Allee, A., Cai, H., Foust, T., & Wyman, C. E., (2017).
Cellulosic ethanol: status and innovation. Current Opinion in Biotechnology, 45, 202–211.
89. Morán, K., Moriarty, K. L., Milbrandt, A. R., Warner, E., Lewis, J. E., & Schwab, A. A.,
(2017). 2016 Bioenergy Industry Status Report (No. NREL/TP-5400-70397). National
Renewable Energy Lab. (NREL), Golden, CO (United States).
90. Stolark, J., (2017). “1.5 Gen Technologies Could Boost Cellulosic Ethanol Production
by Nearly 2 Billion Gallons”. Ideas, Insights, Sustainable, Solutions. Environmental and
Energy Study Institute (EESI).
91. Mueller, K., & Krissek, G., (2014). Future Opportunities and Challenges for Ethanol
Production and Technology.
92. Liu, C. G., Xiao, Y., Xia, X. X., Zhao, X. Q., Peng, L., Srinophakun, P., & Bai, F. W.,
(2019). Cellulosic ethanol production: Progress, challenges and strategies for solutions.
Biotechnology Advances, 37(3), 491–504.
93. Ramos, L. P., Suota, M. J., Fockink, D. H., Pavaneli, G., Da Silva, T. A., & Łukasik, R.
M., (2020). Enzymes and biomass pretreatment. In: Recent Advances in Bioconversion
of Lignocellulose to Biofuels and Value-Added Chemicals within the Biorefinery
Concept (pp. 61–100). Elsevier.
94. Carrillo-Nieves, D., Saldarriaga-Hernandez, S., Gutiérrez-Soto, G., Rostro-Alanis, M.,
Hernández-Luna, C., Alvarez, A. J., & Parra-Saldívar, R., (2020). Biotransformation of
agro-industrial waste to produce lignocellulolytic enzymes and bioethanol with a zero
waste. Biomass Conversion and Biorefinery.
95. De Lucca, M. R. Z., (2020). Empirical Analysis of Production Decision Determinants
of Sugar and Ethanol in the Sugarcane Agroindustry Paulista (Doctoral dissertation).
96. Soccol, C. R., De Souza, V. L. P., Medeiros, A. B. P., Karp, S. G., Buckeridge, M.,
Ramos, L. P., & Da Silva, B. E. P., (2010). Bioethanol from lignocelluloses: Status and
perspectives in Brazil. Bioresource Technology, 101(13), 4820–4825.
97. Lopes, M. L., De Lima, P. S. C., Godoy, A., Cherubin, R. A., Lorenzi, M. S., Giometti,
F. H. C., & De Amorim, H. V., (2016). Ethanol production in Brazil: A bridge between
science and industry. Brazilian Journal of Microbiology, 47, 64–76.
98. UNICA, (2020). Final report 2019/20 crop year in the center-south region of Brazil.
http://unicadata.com.br/listagem.php?idMn=118 (accessed on 28 October 2021).
99. Dias, M. O., Junqueira, T. L., Cavalett, O., Cunha, M. P., Jesus, C. D., Rossell, C. E., &
Bonomi, A., (2012). Integrated versus stand-alone second-generation ethanol production
from sugarcane bagasse and trash. Bioresource Technology, 103(1), 152–161.
100. Quintero, J. A., Montoya, M. I., Sánchez, O. J., Giraldo, O. H., & Cardona, C. A.,
(2008). Fuel ethanol production from sugarcane and corn: Comparative analysis for a
Colombian case. Energy, 33(3), 385–399.
101. Borges, J. R., (2018). Obtenção de bioetanol por fermentação a partir do milho.
Apple Academic Press
102. Paula, C. E. D. E. T., & Biaggi, D. E., (2017). Technological innovations and trends
in the production of second-generation ethanol from sugarcane through the enzymatic
hydrolytic route: A technological prospection study.
103. Pereira, Jr. N., Couto, M. A. P. G., & Santa, A. L. M., (2008). Biomass of lignocellulosic
composition for fuel ethanol production and the context of biorefinery. Series on
Biotechnology, 2, 2–45.
104. Bechara, R., Gomez, A., Saint-Antonin, V., Schweitzer, J. M., & Maréchal, F., (2016).
Methodology for the design and comparison of optimal production configurations of
first and first and second generation ethanol with power. Applied Energy, 184, 247–265.
Author Copy
105. Manochio, C., (2014). Bioethanol production from sugarcane, corn and sugar beet: A
comparison of technological, environmental and economic indicators. Course completion
work (Chemical Engineering)–Federal University of Alfenas. Wells of Caldas, MG.
106. Miguel, J. V. P., (2013). Integrated production of first and second generation sugarcane
bioethanol: energy, environmental and economic analysis (Doctoral dissertation, Univer-
sidade De São Paulo).
107. Lopes, E. M., (2006). Energy analysis and technical feasibility of producing biodiesel
from beef tallow. Itajubá: UNIFEI.
108. PECEGE- Continuing education program in economics and business management,
(2012). Sugarcane, sugar and ethanol production costs in Brazil: 2011/2012 harvest
closing. Piracicaba: University of São Paulo, Superior School of Agriculture “Luiz de
Queiroz,” Continuing Education Program in Business Economics and Management/
Department of Economics, Administration and Sociology. P. 50. Report presented to the
Confederation of Agriculture and Livestock of Brazil (CNA).
109. Donke, A. C. G., et al., (2013). Energo and exergo-environmental analysis of a multipur-
pose process for ethanol production from sugarcane and corn. In: International Exergy,
Life Cycle Assessment, and Sustainability Workshop & Symposium; Nisyros. Anais.
Nisyros: Ecost, 1.305–1.312, 437–448.
110. Silveira, R., (2017). Brazil Opens First Ethanol Plant Fueled 100% with Corn. Canal Rural.
111. Montipó, et al., (2018). Integrated production of second generation ethanol and lactic
acid from stem exploded elephant grass. Bioresource Technology, 249, 1017–1024.
112. Dias, M. O., Da Cunha, M. P., Maciel, F. R., Bonomi, A., Jesus, C. D., & Rossell, C.
E., (2011). Simulation of integrated first and second generation bioethanol production
from sugarcane: Comparison between different biomass pretreatment methods. Journal
of Industrial Microbiology & Biotechnology, 38(8), 955–966.
113. Oliveira, C. M., Cruz, A. J., & Costa, C. B., (2014). Comparison among proposals for
energy integration of processes for 1G/2G ethanol and bioelectricity production. In:
Computer Aided Chemical Engineering (Vol. 33, pp. 1585–1590). Elsevier.
114. Bergmann, J. C., Trichez, D., Sallet, L. P., E Silva, F. C. D. P., & Almeida, J. R., (2018).
Technological advancements in 1G ethanol production and recovery of by-products based
on the biorefinery concept. In: Advances in Sugarcane Biorefinery (pp. 73–95). Elsevier.
115. Sun, X. Z., Fujimoto, S., & Minowa, T., (2013). A comparison of power generation and
ethanol production using sugarcane bagasse from the perspective of mitigating GHG
emissions. Energy Policy, 57, 624–629.
116. Sharma, A., Singh, Y., Ansari, N. A., Pal, A., & Lalhriatpuia, S., (2020b). Experimental
investigation of the behavior of a DI diesel engine fueled with biodiesel/diesel blends
having effect of raw biogas at different operating responses. Fuel, 279, 118460.
117. Araújo, R. M. D., (2019). Food Acquisition Program (2013–2017): Evaluation of the
Apple Academic Press
Author Copy
of the sugar-energy sector in the State of São Paulo, between 2005 and 2009. Nova
Economia, 25(1), 209–224.
121. Bechara, R., Gomez, A., Saint-Antonin, V., Albarelli, J., Ensinas, A., Schweitzer, J. M.,
& Maréchal, F., (2014). Methodology for minimizing the utility consumption of a 2G
ethanol process. Chemical Engineering, 39.
122. Blümmel, M., et. al., (2018). Ammonia Fiber Expansion (AFEX) as spin off technology
from 2nd generation biofuel for upgrading cereal straws and stover’s for livestock feed.
Animal Feed Science and Technology, 236, 178–186.
123. BNDES, (2019). (Vol. 41, pp. 148–208). http://www.bndes.gov.br/SiteBNDES/export/
sites/default/bndespt/Galerias/Arquivos/conhecimento/revista/rev4174 (accessed on 28
October 2021).
124. Borsatto, J. L. Jr., Souza, R. F., Weiss, L., & Dal, V. D. G., (2015). Theoretical essay on the
feasibility of producing ethanol from corn: A possibility for sugar and ethanol mills and
grain producers in the state of Mato Grosso. In: Annals of IV International Symposium on
Project Management, Innovation and Sustainability. Available from: http://www.singep.
org.br/4singep/resultado/319.pdf.102.CorriganME (accessed on 28 October 2021).
125. Chen, E., Song, H., Li, Y., Chen, H., Wang, B., Che, X., & Zhao, S., (2020). Analysis of
aroma components from sugarcane to non-centrifugal cane sugar using GC-O-MS. RSC
Advances, 10(54), 32276–32289.
126. Da Conceição, G. A., Moysés, D. N., Santa, A. L. M. M., & De Castro, A. M., (2018).
Fed-batch strategies for saccharification of pilot-scale mild-acid and alkali pretreated
sugarcane bagasse: Effects of solid loading and surfactant addition. Industrial Crops and
Products, 119, 283–289.
127. De Farias, S. C. E., & Bertucco, A., (2016). Bioethanol from microalgae and cyanobacteria:
A review and technological outlook. Process Biochemistry, 51(11), 1833–1842.
128. Dias, M. O., Junqueira, T. L., Cavalett, O., Pavanello, L. G., Cunha, M. P., Jesus, C. D.,
& Bonomi, A., (2013). Biorefineries for the production of first and second generation
ethanol and electricity from sugarcane. Applied Energy, 109, 72–78.
129. Dias, M. O., Junqueira, T. L., Jesus, C. D., Rossell, C. E., Maciel, F. R., & Bonomi, A.,
(2012b). Improving second generation ethanol production through optimization of first
generation production process from sugarcane. Energy, 43(1), 246–252.
130. Dias, M. O., Junqueira, T. L., Rossell, C. E. V., Maciel, F. R., & Bonomi, A., (2013b).
Evaluation of process configurations for second generation integrated with first generation
bioethanol production from sugarcane. Fuel Processing Technology, 109, 84–89.
131. Dutta, S., & Pal, S., (2014). Promises in direct conversion of cellulose and lignocellulosic
biomass to chemicals and fuels: Combined solvent-nanocatalysis approach for biorefinary.
Biomass and Bioenergy, 62, 182–197.
132. Ernsting, Y. R. M., Stevens, L., Wardle, A. R., & Hall, J. C., (2016). The Renewable Fuel
Apple Academic Press
Author Copy
worth being flexible. Biotechnology for Biofuels, 6(1), 142.
136. Ge, S., Wu, Y., Peng, W., Xia, C., Mei, C., Cai, L., & Tsang, Y. F., (2020). High-pressure
CO2 hydrothermal pretreatment of peanut shells for enzymatic hydrolysis conversion
into glucose. Chemical Engineering Journal, 385, 123949.
137. Gibbons, W. R., & Hughes, S. R., (2009). Integrated biorefineries with engineered
microbes and high-value co-products for profitable biofuels production. In Vitro Cellular
& Developmental Biology-Plant, 45(3), 218–228.
138. Kang, A., & Lee, T. S., (2015). Converting sugars to biofuels: Ethanol and beyond.
Bioengineering, 2(4), 184–203.
139. Kumar, A. K., & Sharma, S., (2017). Recent updates on different methods of pretreatment
of lignocellulosic feedstocks: A review. Bioresources and Bioprocessing, 4(7).
140. Lennartsson, P. R., Erlandsson, P., & Taherzadeh, M. J., (2014). Integration of the
first and second generation bioethanol processes and the importance of by-products.
Bioresource Technology, 165, 3–8.
141. Lin, M., & Gi-Hyung, R., (2014). Effects of thermomechanical extrusion and particle size
reduction on bioconversion rate of corn fiber for ethanol production. Cereal Chemistry,
91(4), 366–373.
142. Liu, W., Wu, R., Wang, B., Hu, Y., Hou, Q., Zhang, P., & Wu, R., (2020). Comparative
study on different pretreatment on enzymatic hydrolysis of corncob residues. Bioresource
Technology, 295, 122244.
143. Luiza, A. A., Rempel, A., Cavanhi, V. A. F., Alves, M., Deamici, K. M., Colla, L. M.,
& Costa, J. A. V., (2020). Simultaneous saccharification and fermentation of Spirulina
sp. and corn starch for the production of bioethanol and obtaining biopeptides with high
antioxidant activity.
144. Mass, R. A. (2018). Utilization of distiller’s dried grains with solubles (DGGS) by cattle.
In: Ingledew, W. M., Austin, G. D., Kelsall, D. R., & Kluhspies, C., (eds.), The Alcohol
Textbook (5th edn.). A reference for the beverage, fuel, and industrial alcohol industries,
Nottingham.
145. Maurya, D. P., Singla, A., & Negi, S., (2015). An overview of key pretreatment
processes for biological conversion of lignocellulosic biomass to bioethanol. 3 Biotech,
5, 597–609.
146. Mendes, F. M., Dias, M. O. S., Ferraz, A., Milagres, A. M. F., Santos, J. C., & Bonomi,
A., (2017). Techno-economic impacts of varied compositional profiles of sugarcane
experimental hybrids on a biorefinery producing sugar, ethanol and electricity. Chemical
Engineering Research and Design, 125, 72–78.
147. Modesto, M., Zemp, R. J., & Nebra, S. A., (2009). Ethanol production from sugarcane:
Assessing the possibilities of improving energy efficiency through exergetic cost
analysis. Heat Transfer Engineering, 30(4), 272–281.
148. Mojović, L., Nikolić, S., Rakin, M., & Vukasinović, M., (2006). Production of bioethanol
Apple Academic Press
Author Copy
second-generation ethanol production. Renewable and Sustainable Energy Reviews,
131, 109991.
152. Ferment Technologies in Sugar and Alcohol Ltd. (2009). Process of reusing yeast
biomass in integration Alcoholic fermentation of sugarcane and starchy substrates (Vol.
103, pp. 323–335). 2015. BR10 2015 021056 6.104. Nottingham University Press.
153. Oliveira, C. M., Cruz, A. J. G., & Costa, C. B. B., (2016). Improving second generation
bioethanol production in sugarcane biorefineries through energy integration. Applied
Thermal Engineering, 109, 819–827.
154. Palacios-Bereche, R., Mosqueira-Salazar, K. J., Modesto, M., Ensinas, A. V., Nebra, S.
A., Serra, L. M., & Lozano, M. A., (2013). Exergetic analysis of the integrated first-and
second-generation ethanol production from sugarcane. Energy, 62, 46–61.
155. Petersen, A. M., Aneke, M. C., & Görgens, J. F., (2014). Techno-economic comparison
of ethanol and electricity coproduction schemes from sugarcane residues at existing
sugar mills in Southern Africa. Biotechnology for Biofuels, 7(1), 105.
156. Rajak, R. C., & Banerjee, R., (2020). An innovative approach of mixed enzymatic
venture for 2G ethanol production from lignocellulosic feedstock. Energy Conversion
and Management, 207(1).
157. Ramos, L. P., Da Silva, L., Ballem, A. C., Pitarelo, A. P., Chiarello, L. M., & Silveira,
M. H. L., (2015). Enzymatic hydrolysis of steam-exploded sugarcane bagasse using high
total solids and low enzyme loadings. Bioresource Technology, 175, 195–202.
158. Rezania, S., Oryani, B., Cho, J., Talaiekhozani, A., Sabbagh, F., Hashemi, B., & Moham-
madi, A. A., (2020). Different pretreatment technologies of lignocellulosic biomass for
bioethanol production: An overview. Energy, 117457.
159. Rezic, T., Oros, D., Markovic, I., Kracher, D., Ludwig, R., & Santek, B., (2013). Integrated
hydrolyzation and fermentation of sugar beet pulp to bioethanol. J Microbiol Biotechnol,
23(9), 1244–1252.
160. Saini, J. K., Saini, R., & Tewari, L., (2015). Lignocellulosic agriculture wastes as biomass
feedstocks for second-generation bioethanol production: Concepts and recent develop-
ments. 3 Biotech, 5(4), 337–353.
161. Santos, F., Eichler, P., De Queiroz, J. H., & Gomes, F., (2020). Production of second-gener-
ation ethanol from sugarcane. In: Sugarcane Biorefinery, Technology and Perspectives (pp.
195–228). Academic Press.
162. Seabra, J. E., & Macedo, I. C., (2011). Comparative analysis for power generation and
ethanol production from sugarcane residual biomass in Brazil. Energy Policy, 39(1),
421–428.
163. Sewsynker-Sukai, Y., & Kana, E. G., (2018). Simultaneous saccharification and
bioethanol production from corn cobs: Process optimization and kinetic studies.
Bioresource Technology, 262, 32–41.
164. Shen, J., & Agblevor, F. A., (2010). Modeling semi-simultaneous saccharification and
Apple Academic Press
Author Copy
167. Testoni, F. M., & Expressão, P. E. C. D., (2017). National Energy Research Center in
Materials – CNPEM National Bioethanol Science and Technology Laboratory-CTBE.
168. Van, Z. W. H., Bloom, M., & Viktor, M. J., (2012). Engineering yeasts for raw starch
conversion. Applied Microbiology and Biotechnology, 95(6), 1377–1388.
169. Vargas, F., Domínguez, E., Vila, C., Rodríguez, A., & Garrote, G., (2015). Agricultural
residue valorization using a hydrothermal process for second generation bioethanol and
oligosaccharides production. Bioresource Technology, 191, 263–270.
170. Zhang, H., Zhang, J., Xie, J., & Qin, Y., (2020). Effects of NaOH-catalyzed organosolv
pretreatment and surfactant on the sugar production from sugarcane bagasse. Bioresource
Technology, 123601.
171. Zhao, J., Xu, Y., Wang, W., Griffin, J., Roozeboom, K., & Wang, D., (2020). Bioconversion
of industrial hemp biomass for bioethanol production: A review. Fuel, 281, 118725
(biomass).
172. Zhao, Y., Damgaard, A., Liu, S., Chang, H., & Christensen, T. H., (2020). Bioethanol from
corn stover-Integrated environmental impacts of alternative biotechnologies. Resources,
Conservation and Recycling, 155, 104652.
173. Lin, C. C., Kang, J. R., Huang, G. L., & Liu, W. Y. (2020). Forest biomass-to-
biofuel factory location problem with multiple objectives considering environmental
uncertainties and social enterprises. Journal of Cleaner Production, 262, 121327.
174. Macrelli, S., Galbe, M., & Wallberg, O. (2014). Effects of production and market factors
on ethanol profitability for an integrated first and second generation ethanol plant using
the whole sugarcane as feedstock. Biotechnology for Biofuels, 7(1), 1–16.
175. Dantas, G. A., Legey, L. F., & Mazzone, A. (2013). Energy from sugarcane bagasse in
Brazil: An assessment of the productivity and cost of different technological routes.
Renewable and Sustainable Energy Reviews, 21, 356–364.
176. Sartana & Afriyeni, N. (2017). Perundungan maya (Cyber Bullying) pada remaja awal.
Jurnal Psikologi Insight, 1(1), 25–39.
177. Markel, E., Sims, C., & English, B. C. (2018). Policy uncertainty and the optimal invest-
ment decisions of second-generation biofuel producers. Energy Economics, 76, 89–100.
178. Gómez, Á & Aristizábal, V., (2010). Biorefineries based on coffee cut-stems and sugarcane
bagasse: Furan-based compounds and alkanes as interesting products. Bioresource
technology, 196, 480–489.
179. García, C. A., Fuentes, A., Hennecke, A., Riegelhaupt, E., Manzini, F., & Masera, O.
(2011). Life-cycle greenhouse gas emissions and energy balances of sugarcane ethanol
production in Mexico. Applied Energy, 88(6), 2088–2097.
180. García, C. A., & Manzini, F. (2012). Environmental and economic feasibility of
sugarcane ethanol for the Mexican transport sector. Solar Energy, 86(4), 1063–1069.
181. Garcia, M. L. et al. (2015). Bioenergy from stillage anaerobic digestion to enhance the
energy balance ratio of ethanol production. Journal of Environmental Management, 162,
Apple Academic Press
102–114.
182. Hennecke, A. M., Mueller-Lindenlauf, M., García, C. A., Fuentes, A., Riegelhaupt, E.,
& Hellweg, S. (2016). Optimizing the water, carbon, and land-use footprint of bioenergy
production in Mexico-Six case studies and the nationwide implications. Biofuels,
Bioproducts and Biorefining, 10(3), 222–239.
183. Shuba, E. S., & Kifle, D. (2018). Microalgae to biofuels:‘Promising’ alternative and
renewable energy, review. Renewable and Sustainable Energy Reviews, 81, 743–755.
184. Renewable Fuels Association (RFA). (2019). World Fuel Ethanol Production. URL
https://ethanolrfa.org/resources/industry/statistics/.
Author Copy
185. IEA, (2018). IEA Agricultural Outlook 2018–2020.
186. BNDES–National Bank for Economic and Social Development. “BNDES the economic
effects of the ethanol first-generation.” BNDES Electronic Portal (2018).
187. CGGE-Bioetanol Combustível: Uma Oportunidade para o Brasil (2009), UNICAMP.
Brasília (2009).
188. García-Olivares, A., Solé, J., & Osychenko, O. (2018). Transportation in a 100% renewable
energy system. Energy Conversion and Management, 158, 266–285.
189. Ramos, J. M., Simó-Alfonso, E. F., Ramis-Ramos, G., Gelfi, C., & Righetti, P. G.
(2000). Determination of cow's milk and ripening time in nonbovine cheese by capillary
electrophoresis of the ethanol-water protein fraction. ELECTROPHORESIS: An
International Journal, 21(3), 633–640.
190. Eijck, J., Romijn, H., Balkema, A., & Faaij, A. (2014). Global experience with jatropha
cultivation for bioenergy: an assessment of socio-economic and environmental aspects.
Renewable and Sustainable Energy Reviews, 32, 869–889.
191. Milanez, D. H., Amaral, R. M. D., Faria, L. I. L. D., & Gregolin, J. A. R. (2013). Assessing
nanocellulose developments using science and technology indicators. Materials Research,
16, 635–641.
192. Schütt, K. T., Arbabzadah, F., Chmiela, S., Müller, K. R., & Tkatchenko, A. (2017).
Quantum-chemical insights from deep tensor neural networks. Nature communications,
8(1), 1–8.
193. Kumar, D., Jansen, M., Basu, R., & Singh, V. (2020). Enhancing ethanol yields in corn
dry grind process by reducing glycerol production. Cereal Chemistry, 97(5), 1026–1036.
194. Pereira, L. G., Cavalett, O., Bonomi, A., Zhang, Y., Warner, E., & Chum, H. L. (2019).
Comparison of biofuel life-cycle GHG emissions assessment tools: The case studies of
ethanol produced from sugarcane, corn, and wheat. Renewable and Sustainable Energy
Reviews, 110, 1–12.
195. Manochio, C., Andrade, B. R., Rodriguez, R. P., & Moraes, B. S. (2017). Ethanol from
biomass: A comparative overview. Renewable and Sustainable Energy Reviews, 80,
743–755.
196. Pachón, E. R., Mandade, P., & Gnansounou, E. (2020). Conversion of vine shoots
into bioethanol and chemicals: Prospective LCA of biorefinery concept. Bioresource
technology, 303, 122946.
197. FAO, (2018). FAO Agricultural Outlook 2018–2020. https://www.fao.org/3/BT092e/
BT092e.pdf.
198. Morandi, F., Perrin, A., & Østergård, H. (2016). Miscanthus as energy crop:
Environmental assessment of a miscanthus biomass production case study in France.
Journal of Cleaner Production, 137, 313–321.
199. Zhan, C., Feng, Z., Zhang, M., Tang, C., & Huang, Z. (2016). Experimental investigation
Apple Academic Press
on effect of ethanol and di-ethyl ether addition on the spray characteristics of diesel/
biodiesel blends under high injection pressure. Fuel, 218, 1–11.
200. Xuan, Z., Naimi, T. S., Kaplan, M. S., Bagge, C. L., Few, L. R., Maisto, S., … &
Freeman, R. (2010). Alcohol policies and suicide: a review of the literature. Alcoholism:
clinical and experimental research, 40(10), 2043–2055.
201. Mojović, L., Nikolić, S., Rakin, M., & Vukasinović, M. (2006). Production of bioethanol
from corn meal hydrolyzates. Fuel, 85(12–13), 1750–1755.
202. Agarwal, A., Rana, M., Qiu, E., AlBunni, H., Bui, A. D., & Henkel, R. (2018). Role
of oxidative stress, infection and inflammation in male infertility. Andrologia, 50(11),
Author Copy
e13126.
203. Cripwell, R. A., Favaro, L., Viljoen-Bloom, M., & van Zyl, W. H. (2020). Consolidated
bioprocessing of raw starch to ethanol by Saccharomyces cerevisiae: Achievements and
challenges. Biotechnology Advances, 42, 107579.
204. Shen, J., & Agblevor, F. A. (2010). Modeling semi-simultaneous saccharification and
fermentation of ethanol production from cellulose. Biomass and Bioenergy, 34(8),
1098–1107.
205. Kurambhatti, C., Kumar, D., & Singh, V. (2019). Impact of fractionation process on the
technical and economic viability of corn dry grind ethanol process. Processes, 7(9), 578.
206. Sadhukhan, J., Martinez-Hernandez, E., Amezcua-Allieri, M. A., & Aburto, J. (2019).
Economic and environmental impact evaluation of various biomass feedstock for
bioethanol production and correlations to lignocellulosic composition. Bioresource
Technology Reports, 7, 100230.
207. Pachón, E. R., Mandade, P., & Gnansounou, E. (2020). Conversion of vine shoots
into bioethanol and chemicals: Prospective LCA of biorefinery concept. Bioresource
Technology, 303, 122946.
208. Peterson, V. L., McCool, B. A., & Hamilton, D. A. (2015). Effects of ethanol exposure
and withdrawal on dendritic morphology and spine density in the nucleus accumbens
core and shell. Brain research, 1594, 125–135.
209. Ramos, W. B., Figueiredo, M. F., Brito, K. D., Ciannella, S., Vasconcelos, L. G., &
Brito, R. P. (2016). Effect of solvent content and heat integration on the controllability
of extractive distillation process for anhydrous ethanol production. Industrial &
Engineering Chemistry Research, 55(43), 11315–11328.
210. Mandira, M., Rampure, M., Todkar, A., & Sharma, P. (2019). Ethanol from maize: an
entrepreneurial opportunity in agrobusiness. Biofuels, 10(3), 385–391.
211. Rezic, T., Oros, D., Markovic, I., Kracher, D., Ludwig, R., & Santek, B. (2013). Integrated
hydrolyzation and fermentation of sugar beet pulp to bioethanol. Journal of Microbiology
and Biotechnology, 23(9), 1244–1252.
212. Cheah, W. Y., Sankaran, R., Show, P. L., Ibrahim, T. N. B. T., Chew, K. W., Culaba, A., &
Jo-Shu, C. (2020). Pretreatment methods for lignocellulosic biofuels production: current
advances, challenges and future prospects. Biofuel Research Journal, 7(1), 1115.
213. Umar, A., Khan, M. A., Kumar, R., & Algarni, H. (2017). Ag-doped ZnO nanoparticles for
enhanced ethanol gas sensing application. Journal of nanoscience and nanotechnology,
18(5), 3557–3562.
214. Gomes, D. G., Serna-Loaiza, S., Cardona, C. A., Gama, M., & Domingues, L. (2018).
Insights into the economic viability of cellulases recycling on bioethanol production from
recycled paper sludge. Bioresource technology, 267, 347–355.
215. Kim, N. J., Li, H., Jung, K., Chang, H. N., & Lee, P. C. (2011). Ethanol production
Apple Academic Press
from marine algal hydrolysates using Escherichia coli KO11. Bioresource technology,
102(16), 7466–7469.
216. Baruah, J., Nath, B. K., Sharma, R., Kumar, S., Deka, R. C., Baruah, D. C., & Kalita,
E. (2018). Recent trends in the pretreatment of lignocellulosic biomass for value-added
products. Frontiers in Energy Research, 6, 141.
217. Waghmare, P. R., Khandare, R. V., Jeon, B. H., & Govindwar, S. P. (2018). Enzymatic
hydrolysis of biologically pretreated sorghum husk for bioethanol production. Biofuel
Res. J, 5(3), 846–853.
218. Nielsen, F., Galbe, M., Zacchi, G., & Wallberg, O. (2020). The effect of mixed agricultural
Author Copy
feedstocks on steam pretreatment, enzymatic hydrolysis, and cofermentation in the
lignocellulose-to-ethanol process. Biomass Conversion and Biorefinery, 10(2), 253–266.
219. Nogueira, K. M. V., Mendes, V., Carraro, C. B., Taveira, I. C., Oshiquiri, L. H., Gupta,
V. K., & Silva, R. N. (2020). Sugar transporters from industrial fungi: key to improving
second-generation ethanol production. Renewable and Sustainable Energy Reviews, 131,
109991.
220. Hilares, R. T., de Almeida, G. F., Ahmed, M. A., Antunes, F. A., da Silva, S. S., Han, J. I.,
& Dos Santos, J. C. (2017). Hydrodynamic cavitation as an efficient pretreatment method
for lignocellulosic biomass: A parametric study. Bioresource technology, 235, 301–308.
221. Enerting, R. M., Shonnard, D. R., Griffing, E. M., Lai, A., & Palou-Rivera, I. (2016). Life
cycle assessments of ethanol production via gas fermentation: anticipated greenhouse gas
emissions for cellulosic and waste gas feedstocks. Industrial & Engineering Chemistry
Research, 55(12), 3253–3261.
222. Bondancia, T. J., Côrrea, L. J., Cruz, A. J., BADINO, A., Mattoso, L. H. C., Marconcini,
J. M., & Farinas, C. S. (2017). Nanofibras de celulose via hidrólise enzimática em reator
de tanque agitado. In Embrapa Instrumentação-Artigo em anais de congresso (ALICE).
In: Workshop Da Rede De Nanotecnologia Aplicada Ao Agronegócio, 9, 2017, Editors:
Caue Ribeiro de Oliveira, Elaine Cristina Paris, Luiz Henrique Capparelli Mattoso,
Marcelo Porto Bemquerer, Maria Alice Martins, Odílio Benedito Garrido de Assis. São
Carlos. Anais. São Carlos: Embrapa Instrumentação, 2017. pp. 287–290.
223. Moran, P. J., Vacek, A. T., Racelis, A. E., Pratt, P. D., & Goolsby, J. A. (2017). Impact
of the arundo wasp, Tetramesa romana (Hymenoptera: Eurytomidae), on biomass of the
invasive weed, Arundo donax (Poaceae: Arundinoideae), and on revegetation of riparian
habitat along the Rio Grande in Texas. Biocontrol Science and Technology, 27(1), 96–114.
224. Patni, N., Pillai, S. G., & Dwivedi, A. H. (2013). Wheat as a promising substitute of corn
for bioethanol production. Procedia Engineering, 51, 355–362.
225. Gupta, R., & Lee, Y. Y. (2010). Investigation of biomass degradation mechanism in
pretreatment of switchgrass by aqueous ammonia and sodium hydroxide. Bioresource
technology, 101(21), 8185–8191.
226. RFA Agricultural Outlook 2016–2020. https:// https://ethanolrfa.org/
227. EMBRAPA, 2016: Faleiro, F. G., & Junqueira, N. T. V. (2016). Maracujá: o produtor
pergunta, a Embrapa responde. Embrapa Cerrados-Livro técnico (INFOTECA-E).
228. Pereira, J. P., Oliveira, V. B., & Pinto, A. M. F. R. (2017). Modeling of passive direct
ethanol fuel cells. Energy, 133, 652–665.
229. Dias, M. O. S., Ferraz, A., Milagres, A. M. F., Santos, J. C., & Bonomi, A. (2017).
Techno-economic impacts of varied compositional profiles of sugarcane experimental
hybrids on a biorefinery producing sugar, ethanol and electricity. Chemical Engineering
Research and Design, 125, 72–78.
Apple Academic Press
230. Pereira, L. G., Dias, M. O., Junqueira, T. L., Pavanello, L. G., Chagas, M. F., Cavalett,
O., … & Bonomi, A. (2014). Butanol production in a sugarcane biorefinery using
ethanol as feedstock. Part II: Integration to a second generation sugarcane distillery.
Chemical Engineering Research and Design, 92(8), 1452–1462.
231. DeLuca, T. H., Gundale, M. J., Brimmer, R. J., & Gao, S. (2020). Pyrogenic carbon
generation from fire and forest restoration treatments. Frontiers in Forests and Global
Change, 3, 24.
232. Chen, Z., Jacoby, W. A., & Wan, C. (2019). Ternary deep eutectic solvents for effective
biomass deconstruction at high solids and low enzyme loadings. Bioresource technology,
Author Copy
279, 281–286.
233. Schuette, S. E. (2017). Insights into thermophilic plant biomass hydrolysis from
Caldicellulosiruptor systems biology. Microorganisms, 8(3), 385.
234. Gibbens, W., & Hughes, S. (2009). Integrated biorefineries with engineered microbes
and high-value co-products for profitable biofuels production. Biofuels, 265–283.
235. Carvalho, D. J., Moretti, R. R., Colodette, J. L., & Bizzo, W. A. (2020). Assessment
of the self-sustained energy generation of an integrated first and second generation
ethanol production from sugarcane through the characterization of the hydrolysis
process residues. Energy Conversion and Management, 203, 112267. https://www.
conab.gov.br/.
236. Gissén, C., Prade, T., Kreuger, E., Nges, I. A., Rosenqvist, H., Svensson, S. E., &
Björnsson, L. (2014). Comparing energy crops for biogas production–Yields, energy input
and costs in cultivation using digestate and mineral fertilisation. Biomass and bioenergy,
64, 199–210.
237. Furlan, A. L., Bianucci, E., del Carmen Tordable, M., Castro, S., & Dietz, K. J. (2014).
Antioxidant enzyme activities and gene expression patterns in peanut nodules during a
drought and rehydration cycle. Functional Plant Biology, 41(7), 704–713.
238. Peterson, T., Sharma, N., Shojaeiarani, J., & Bajwa, S. G. (2014). A review of densified
solid biomass for energy production. Renewable and Sustainable Energy Reviews, 96,
296–305.
239. Kazi Kazeema, N., Borana Pravin, H., & Ikale Vijay, H. (2021). Formulation and
Evaluation of An Anti-Oxidant Product From Betel Leaf Extract.
240. Kathi, Y., An, M., Liu, K., Nagai, S., Shigematsu, T., Morimura, S., & Kida, K. (2006).
Ethanol production from acid hydrolysate of wood biomass using the flocculating yeast
Saccharomyces cerevisiae strain KF-7. Process Biochemistry, 41(4), 909–914.
241. Werner, F. (2017). Background report for the life cycle inventories of wood and wood
based products for updates of ecoinvent 2.2.
242. Dias, M. O., Lima, D. R., & Mariano, A. P. (2018). Techno-economic analysis of
cogeneration of heat and electricity and second-generation ethanol production from
sugarcane. In Advances in Sugarcane Biorefinery (pp. 197–212). Elsevier.
243. Patrizi, N., Pulselli, F. M., Morandi, F., & Bastianoni, S. (2010). Evaluation of the
emergy investment needed for bioethanol production in a biorefinery using residual
resources and energy. Journal of Cleaner Production, 96, 549–556.
244. Šobotník, J., Bourguignon, T., Carrijo, T. F., Bordereau, C., Robert, A., Křížková, B.,
& Cancello, E. M. (2015). The nasus gland: a new gland in soldiers of Angularitermes
(Termitidae, Nasutitermitinae). Arthropod Structure and Development, 44(5), 401–406.
245. De Lucca, M. R. Z. (2020). Análise empírica sobre os determinantes da decisão de
Apple Academic Press
Author Copy
Author Copy
1
University of Nuevo León, Faculty of Agronomy, Francisco Villa S/N Col.
Ex Hacienda El Canadá – 66415, General Escobedo, N. L., México,
E-mail: francisco.zavala.garcia@gmail.com (F. Zavala-García)
2
Biorefinery Group, Food Research Department,
Faculty of Chemistry Sciences, Autonomous University of Coahuila,
Saltillo–25280, Coahuila, Mexico
ABSTRACT
The continuous decrease in fossil fuel reserves and their impact on the
environment, lead the search for new alternative energy sources. In this
context, biofuels provide a viable substitute that contributes to the reduction
of polluting emissions into the atmosphere and increases the conventional
fuels efficiency. However, biofuels confront great challenges such as
obtaining raw material, requiring biomass sources to mitigate this problem.
In this sense, the sweet sorghum variety ROGER has shown potential in the
production of bioethanol, which depending on the production conditions can
have ethanol concentrations in the juice of 81.2 g/l. This chapter mentions
aspects related to the bioethanol production from sweet sorghum juice.
11.1 INTRODUCTION
The continuous use of fossil fuels has led to the reduction of their reserves and
environmental pollution because of the emission of considerable volumes
of GHG and other pollutants into the atmosphere [1]. However, over the
The biofuels production faces four major challenges: 1) the raw material
must not compete with food production, 2) it must have a neutral balance of
polluting emissions (CO2), 3) its production must be continuous, ensuring the
supply of industrial facilities (extraction plants, biorefineries, fermenters),
and 4) extraction methods must be profitable from the energy and economic
point of view to compete in the long term with the price of other sources of
non-renewable energy [5].
Author Copy
Bioenergy is the most extensively used energy source in the world,
mainly in solid form and to a minor extent in liquid and gaseous form. Thus,
the continuous technological advance has allowed the development of more
efficient biomass conversion technologies that make it possible to obtain
biofuels in their different physical states. However, most of these achieve-
ments have not been implemented in commercial facilities yet [6, 7].
With respect to liquid biofuels (bioethanol and biodiesel), they are used
as mixtures or substitutes for gasoline and diesel, being used mainly in
transportation and industry. According to the origin of the raw material and
the technology used to obtain it, they are classified as first, second, third, and
fourth generation, the latter still under development [5, 8], which depend
on sources of sugars, starch, and LCB. Where the conversion to bioethanol
depends on the nature of the raw material, mainly on its biochemical
composition [7].
In the case of first-generation bioethanol, it is a product that derives
mainly from the sugars fermentation contained in agricultural crops with a
high sucrose content, such as sugar cane, beet, and sweet sorghum, or from
the starch contained in the corn grain, wheat, barley, tubers (like potato) and
roots (like cassava), with the disadvantage of being crops intended for food,
so their use as energy material generates a food safety controversy. In addi-
tion to contributing to deforestation by eliminating large areas of primary
forests for their establishment [9]. However, sweet sorghum [Sorghum
bicolor (L.) Moench] has agronomic characteristics that allow it to supply
raw material for the production of biofuels, without affecting the diet and
agricultural areas [10, 11], since it is the food base of millions of people,
mainly in semi-arid regions around the world [12].
Sweet sorghum is a short-cycle plant with a high degree of tolerance
to biotic and abiotic stresses, which can achieve high biomass yields
Author Copy
[19]. They have a high influence on the development of the crop, on the energy
consumptions of the system [20] and the definition of the optimal moment of
its harvest [17]. Although this would not be complete without the study of the
fermentation kinetics that defines the final production of bioethanol [21]. For
this reason, ethanol productivity varies between sweet sorghum genotypes,
due to its characteristics such as the amount of juice they can retain in their
stems and the concentration of sugars, which can influence the process of
conversion to bioethanol, where the temperature and the size of the yeast
inoculum play an important role. The latest has a great influence on the typical
biomass conversion processes, specifically in the sugar fermentation, where
one of the traditionally used yeasts is Saccharomyces cerevisiae [7].
Thus, the potential of sorghum in the bioethanol production is also
determined by the energy necessary to obtain it, energy balances being
necessary in order to reduce consumption and achieve high energy efficiency.
In this sense, an agricultural operation is efficient when the energy obtained
is greater than that used. However, the energy ratio depends on the crop,
the production system and the intensity of management. The last is mainly
associated with the optimal management of the resources used in the
production process such as water, fuels, fertilizers, pesticides, machinery,
electricity, and others [22].
In general, obtaining bioethanol is a complex process that must be
continuously improved, even more so by the introduction of new raw
materials. Therefore, the incorporation of a new plant material always requires
the study and evaluation of its potential application, as well as the study of
the sweet sorghum variety. This chapter give general aspects related to the
production of bioethanol from sweet sorghum juice, considering agronomic
and energetic parameters of the crop, besides of the conversion process to
bioethanol, using a new sweet sorghum variety developed in the Universidad
Author Copy
24]. The morphological and physiological variations of sorghum plants have
provided the material for natural selection to act on them and accumulate
individual differences in a similar or artificial way. These differences allow
the different uses of cultivated sorghums, such as the production of ethanol,
grain, honey, brooms, etc. In such a way, that they are currently grouped
into various categories, based on their main products and uses, as well as the
distinctive or genetic characteristics of the plant [24, 25].
Taxonomically, sorghum was first described by Linnaeus in 1753 under
the name Holcus. He originally delineated several species of Holcus, some of
which were later moved to the Avenae tribe, where the generic name Holcus
now belongs. In 1794, Moench distinguished the genus Sorghum from
the genus Holcus [26]. The genus Sorgum is divided into five subgenera:
sorghum, chaetosorghum, heterosorghum, parasorghum, and stiposorghum.
Within the Sorghum subgenus, the wild species, S. bicolor, S. halepense, and
S. propinquum can be mentioned [27].
Sorghum exhibits several morphophysiological forms and great variation
for flower morphology, resulting in the classification of several basic and
intermediate races. Harlan and de Wet [96], classified Sorghum bicolor (L.)
Moench, subspp. bicolor in five basic races and ten hybrids as shown in the
following subsections.
Author Copy
• Race kafir-durra (KD).
The different uses of sorghum and its high capacity for adaptation to
different climatic conditions have maintained its productive levels, as well
as the areas devoted to this crop. The global trend of sorghum production is
shown in Table 11.1. According to Mundia et al. [28], there are mainly 10
factors that significantly impact sorghum production: (1) climate change,
(2) population growth/economic development, (3) non-food demand, (4)
agricultural inputs, (5) demand for other crops, (6) scarcity of agricultural
resources, (7) biodiversity, (8) cultural influence, (9) prices, and (10) armed
conflicts. Furthermore, several of these factors can simultaneously affect
sorghum production and their degree of impact can be the combination of
several factors, and their magnitude differs geographically.
Author Copy
• Sugar Drip: High productivity with late sowing, susceptible to most
sorghum diseases, with approximately 110 days to maturity (early
maturity).
• M81-E: Late maturing, resistant to lodging and diseases.
• Theis and Brandes: Late maturing, at least 2–3 weeks later than
Dale, resistant to lodging and disease.
TABLE 11.2 Agronomic Performance of Some Sweet Sorghum Varieties Used to Produce
Bioethanol
Varieties Juice (m3/ha) °Bx (%) Ethanol (m3/ha) References
Keller 18.9–20.8 16.5–18.5 1.487–1.544 [31, 32]
Dale 21.1–22.9 15.3–18.7 1.331–1.360 [31, 32]
Della 18.2–24.2 13.7–16.2 1.281–1.429 [31, 32]
M81-E 23.1–23.4 14.9–17.1 1.419–1.496 [31, 32]
Sugar Drip 9.7–13.8 14.4–16.3 0.704–0.842 [31, 32]
Author Copy
11.6 SWEET SORGHUM PRODUCTION
Sweet sorghum is one of the energy crops with high potential for the
bioethanol production, which can be obtained from the juice of the stems,
grains, bagasse, and straw. However, the cultivars with greater interest stand
out in certain agronomic traits such as: (i) high biomass yield; (ii) stems
rich in juices with high sugar content; (iii) high percentage of extractable
juice; (iv) stem geometry capable of resisting lodging; (v) a long period of
industrial use that prolongs the harvest season; and (vi) tolerance to different
production conditions [11, 30].
Compared to other plant materials with potential in the bioethanol produc-
tion such as sugar cane, sugar beet, and corn, sweet sorghum has higher
levels of directly fermentable reducing sugars in its stems and requires fewer
inputs per production unit of biomass. In addition, it has high adaptability to
adverse edaphoclimatic conditions and a greater efficiency in the use of solar
radiation and nitrogen fertilizers than corn and sugar cane [34, 37].
Although sweet sorghum is a crop with a great capacity to adapt to
various agroclimatic conditions, crop yield in any region is associated with
plant growth, which depends on climatic, biological, edaphic factors and
agronomic practices. Some of these factors are summarized below:
1. Precipitation and Temperatures: It adapts to a wide range of
annual precipitations (from 550 to 800 mm) and temperatures (15
to 45°C), under various climatic and soil conditions. Although the
optimum temperature for plant development is between 25 and 40°C
and requires a day duration between 10–14 h [38]. However, these
Author Copy
the initial soil conditions to favor the growth of the crop and increase
yield [41, 97]. Ramírez et al. [42] reported that in Mexico there are
more than 19 million hectares with high optimal conditions under
irrigation conditions, although, with the supply of water, it can be
cultivated throughout the country.
3. Seedtime: Sweet sorghum is an annual plant with a short life cycle
(approximately 4 months) that, depending mainly on the agrocli-
matic conditions and the genotype, can be planted two or three times
a year [14, 43]. According to Han et al. [44], sweet sorghum can
be planted from mid-March to early June with favorable results.
Nevertheless, planting sorghum in early. May and harvested at the
hard dough stage can substantially increase the fermentable sugar
content. Though, a late seedtime can decrease the efficiency of the
use of solar radiation and consequently the productivity of the crop
[45], which may be due to the gradual decrease of the foliar area and
the height of the plant as a consequence of sensitivity to photoperiod
or shorter days [46]. In addition, it delays the harvest exposing the
plants to the attack of pests and diseases that are predominant at the
end of the crop cycle [47]. In general, the sowing date can influence
the contents of sucrose, total sugar, pH, electrical conductivity of the
juice and the Brix degrees [48].
4. Sowing Density: It influences the growth parameters of sorghum and
its yield. For example, optimal plant density can increase biomass
yields, since they maintain a high and stable net assimilation rate,
besides a positive balance between photosynthesis and respiration
processes [40, 49, 50]. Since the plants have a greater vertical angle
of the leaf with respect to the main stem, increasing the capture of
light and consequently a higher photosynthetic rate [51, 52]. In this
sense, it has been observed that a high planting density (≈16.7 plants/
m2) increases the juice, sugar, and biomass yields. However, the stem
diameter and the internodes number decrease contributing to a higher
lodging rate [49, 53].
Apple Academic Press
Author Copy
practices that maintain the necessary levels of N and C in the soil
[56]. Thus, the physical-chemical analysis of the soil is necessary
to make up for its deficiencies. In this sense, a positive response of
sweet sorghum has been observed with applications of ≥40 kg/ha of
N, increasing the biomass, juice, and sugars yields [57–60].
6. Pests Management and Control, Diseases, and Weeds: Undesir-
able plants compete for nutrients, water, and light, they can release
substances through their roots and leaves that turn out to be toxic to
crops. In addition, they provide a favorable habitat for the proliferation
of other pests (arthropods, mites, pathogens, and others) by serving as
their hosts, which can affect the harvesting process and contaminate the
production obtained [61]. On the other hand, pest, and disease cause
considerable crop losses; for example, the yellow aphid Melanaphis
sacchari causes crop damage at all stages of sorghum development [62,
63]. This pest feeds on the sap of the plant, reducing its vigor and yield
[64], and even causing 100% of crop losses under severe attack [65].
7. Irrigation: It tries to supply rainwater when it is insufficient to
supply the water needs of the crop, maintaining productive soils with
optimal humidity levels for growth, in addition to contributing to
increased agricultural yields. Water helps plants in the absorption of
nutrients from the soil and to perform other physiological functions
[66]. Sweet sorghum is a drought-resistant crop that, compared to
other crops (such as sugar cane), requires a lower volume of irrigation
water (approximately 400 mm), which depends mainly on climatic
conditions and the variety [67, 68]. Thus, there is the possibility of
reducing irrigation water needs through sweet sorghum cultivars
with deep roots that intercept most rainwater and in some productive
environments no require irrigation [69].
sorghum harvest may be carried out from the soft mass stage of
the grain (the filling of the grains is complete and they begin to
harden) and extend to physiological maturity (the grains begin to
dry out), which would imply between 104 to 130 days after sowing.
Although, this time will depend on the variety and environmental
conditions [71–74].
Once harvested, sweet sorghum faces some challenges such as
Author Copy
the rapid fermentation of sugars, generating an acceleration of its
decomposition in a few hours, indicating that it should not be stored
for long periods [70, 75]. Therefore, the rapid stem grinding and the
processing of the extracted juice (4 to 5 hours) is essential considering
the high content of fermentable sugars and the metabolic activity of
contaminating bacteria. However, to reduce this postharvest effect,
the juice may be refrigerated, or chemical preservatives can be added
[76, 77]. However, it could change the chemical composition of the
juice [78].
In general, the bioethanol production requires the continuous
supply of sweet sorghum, the transport and storage of high volumes
of plant material, as well as minimal post-harvest losses of ferment-
able sugars [34].
The energy balance allows evaluating the efficiency of the crop production
methods, considering the energy inputs and outputs of the production system.
The energy balance differs widely between agricultural production systems
and life cycle analysis is commonly used to assess the energy efficiency and
environmental effect of bioethanol production. Therefore, energy demand
is classified according to the energy inputs of each production system
[79]. Accordingly, an efficient process requires that the energy invested in
producing a unit of biofuel, including the agricultural and industrial stages,
be less than the energy leaving the system.
In the sweet sorghum production, the inputs associated with agricultural
operations are decisive in energy consumption. Wherein production systems
with low input requirements are characterized by lower consumption and,
hence, greater energy efficiency. In this sense, it has been reported that the
greatest contribution to energy consumption corresponds to the diesel used
to move agricultural machinery and in nitrogen fertilization [22, 80]. There-
upon, consumption of energy has a wide range of variation depending on the
Apple Academic Press
Author Copy
without tillage and without fertilizer with an energy consumption of 2,260
MJ/ha. For the bioethanol production, an energy production of 22.6 to 70.5
GJ/ha has been obtained from fresh sweet sorghum biomass [81]. Likewise,
a net energy gain of 8.37 MJ/l of bioethanol and an energy efficiency of 1.56
have been reported [79].
The juice of the sweet sorghum stalk has shown to be a raw material of
great potential for the biofuels production which, after being transformed
(Figure 11.1) into anhydrous ethanol, is mixed with gasoline and is used in
motors. During the bioethanol production, the stems are crushed to extract
the juice rich in fermentable sugars with an exact proportion that varies
between genotypes (53–85% sucrose, 9–33% glucose and 6–21% fructose)
and that can be converted directly into ethanol with high efficiency, mainly
by the yeast Saccharomyces cerevisiae. Which is a fast-acting microor-
ganism that tolerates high alcohol concentrations (up to 150 g/l) and shows
high bioethanol yields. Furthermore, it maintains high cell viability under
different fermentation conditions [30, 82, 83]. When used in sweet sorghum
juice in the bioethanol production, its growth curve shows three phases: the
exponential phase observed approximately in the first 18 h of incubation, the
slow phase observed approximately between 18 to 24 h and the stationary
phase observed approximately between 24 to 72 hours [95]. Likewise, the
kinetics of bioethanol production depend on several factors such as tempera-
ture, yeast load, pH, aeration time and rate, stirring conditions, nitrogen
source and initial sugar concentration. Therefore, in the bioethanol produc-
tion from sweet sorghum juice, a wide range of yield between 39.2 and 127.8
g/l has been reported [83–85].
Author Copy
FIGURE 11.1 Conventional process diagram for obtaining first-generation bioethanol from
sweet sorghum juice.
The variety ROGER was registered under the same name and with
registration number SOG-261-050315 by the Universidad Autónoma de
Nuevo León in the National Catalog of Plant Varieties. This genotype
is characterized by an average of 75 days to flowering and a cultivation
cycle of 130 days. Depending on the production conditions, this crop have
approximately a fresh biomass yield of 51.66 t/ha, juice of 22.53 m3/ha and
°Bx of 16.04% [74].
ROGER was produced in the Marín experimental field of the Faculty
of Agronomy belonging to the Universidad Autónoma de Nuevo León
(UANL), located in the municipality of Marín, in the state of Nuevo Leon,
Mexico. This is geographically located at 25° 52´ 13.5´´ north latitude
and 100° 02´ 22.56´´ west longitude, with an altitude of 355 meters above
sea level. The climate corresponds to a BS1 (hˊ) hw (é), described as a dry
hot steppe with rains in summer, annual rainfall that fluctuates between
250–500 mm and average annual temperature of 22°C. The type of soil is
vertisols, thin, compacted, with high content of clay and calcium carbonate,
low content of organic matter and pH between 7.5–8.5. ROGER, like other
Author Copy
depth of 0.05 m and distance between rows of 0.8 m, using an experimental
planter (Almaco® CTS EO, Nevada, Iowa, USA). Organic fertilization
(chicken manure) was applied 20 days before sowing with doses of 3 t/ha
(N-60 kg/ha, P-65 kg/ha and K-75 kg/ha) and 20 days after sowing thinning
was performed to obtain an average density of 18 plants/m2. For pest control,
a dose of 0.6 l/ha of the pesticide Pounce 340 CE® was applied, 28 and 59
days after sowing and in particular for the yellow aphid a dose of 0.3 l/ha of
Murralla® was applied at 32 and 81 days after sowing. For weed control, 1 l/
ha of Phyto Amina 40® was applied at 35 and 64 days after sowing. Irrigation
by band (superficial drip) was carried out before sowing and five auxiliary
irrigation after sowing at intervals of 14 days. The harvest was carried out
manually (130 days after sowing) at a cutting height of 0.03 m to 0.04 m
from the stem base. However, to determine the harvest moment with the
highest bioethanol yield, samples were taken at different plant physiological
stages (PS) described by Vanderlip and Reeves [86]. In the first sampling, the
grains showed soft mass (PS7). In the second sampling, the grains showed a
stage of dough (PS8). In the third sampling, the culture was at physiological
maturity (PS9).
In our study, the tillage and fertilization effect on energy consumption during
the establishment of the crop was evaluated. The tillage treatments consisted
of a minimum tillage system, conventional tillage and conventional tillage
with breaking of the plow layer; while the fertilization treatments consisted
of the use of organic and inorganic fertilizer and without fertilizer. For the
above, the methodology described by De las Cuevas et al. [87] and Paneque
and Sánchez [88], based on the proposals of Bridges and Smith [89] and
pesticides, and seeds. As a result, it was obtained that the energy demand
varied depending on the production system used, where the systems with
low inputs required less energy supply. In this sense, minimum tillage and
no fertilization showed the lowest energy consumption values with 21.3 and
17.2 GJ/ha, respectively. Likewise, these systems had the highest energy
efficiency values with 15.1 and 18.7, respectively. On the other hand, the
highest net energy production was 340.3 and 351.4 GJ/ha, obtained with
Author Copy
deep conventional tillage and organic fertilization, respectively. Where, the
highest contribution to energy consumption was using diesel used in agricul-
tural machinery in a range of 0.41 to 1.7 GJ/ha and in inorganic fertilization
with 8.4 GJ/ha [74].
the juice to be preserved for long periods without affecting its quality and
consequently the bioethanol yield.
Apple Academic Press
TABLE 11.3 Bioethanol Production Conditions of with Juice of the Sweet Sorghum Variety
ROGER
Treatments Culture Conditions Bioethanol Yield (g/l) Time (h)
Fresh juice Agitation 150 rpm, 2% inoculum 78.1 48
Fresh juice Agitation 150 rpm, 3% inoculum 80.9 48
Fresh juice Agitation 150 rpm, 1% inoculum 52.3 24
Fresh juice Without agitation, 2% inoculum 34.7 24
Author Copy
Frozen juice Agitation 150 rpm, 1% inoculum 47.3 24
Juice PS7*
Agitation 150 rpm, 2% inoculum 70.3 48
Juice PS8* 65.7 48
Juice PS9* 57.9 48
*
Phenological stage of the plant described by Vanderlip and Reeves [86]. All the tests were
carried out in triplicate, with a standard deviation of less than 5%.
In a test with 100 l of juice, the bioethanol production kinetic was deter-
mined. The juice was placed and sterilized in a stainless-steel tank with a
capacity of 200 l and the changes in pH, °Bx, reducing sugar concentra-
tion (RS), and bioethanol concentration were determined at 0, 24, 48 and
72 h of fermentation. The pH measurement was made with a potentiometer
(Corning® pH meter 430, USA). The °Bx were determined with a portable
digital refractometer (ATAGO® 3810, PAL-1, Atago USA Inc., Bellevue,
USA). The RS quantification was performed by the method described by
Miller [93], using glucose as standard.
As a result, the bioethanol production curve showed a logarithmic phase
up to 48 h, maintaining the stationary phase until 72 h (see Figure 11.2),
during which the maximum bioethanol concentration was reached (81.2 g/l).
The concentration of reducing sugars and °Bx showed a decrease of 64.5
and 59.1%, respectively at 72 h. However, the highest consumption rate was
observed at 48 h with 59.6 and 52.3%, respectively. It is worth mentioning
that the juice was not supplemented with other sources of carbon, nitrogen
or other types of nutrients, nor was the pH adjusted during the fermentation
process. Therefore, the behavior of the production kinetics suggests that the
juice obtained from the variety does not have levels of compound that inhibit
the development of yeast.
Author Copy
Bioethanol production kinetic.
FIGURE 11.2
11.10 CONCLUSIONS
Sweet sorghum is a low-input energy crop that has high adaptability to various
production conditions and great potential in bioethanol production. In this
Apple Academic Press
sense, the sweet sorghum variety ROGER has potential in the first-generation
bioethanol production and can be cultivated in semi-arid climates, achieving
a high production of biomass, juice, and sugar. On the other hand, during
the crop establishment, energy consumption oscillates depending on the
production system used, with low-input systems being those with the lowest
energy consumption. The higher energy input was from the consumption of
diesel fuel (0.41 to 1.7 GJ/ha) and of chemical fertilizer (8.4 GJ/ha). Thus,
Author Copy
the higher energy efficiency with values of 15.1 and 18.7 was obtained with
minimum tillage and without fertilizer application, respectively. Whereas for
the stem juice, it does not require supplementation or addition of nutrients
for its conversion to bioethanol. However, the yields fluctuate depending on
the fermentation conditions, where the maximum yield can be obtained after
72 h of fermentation (81.2 g/l).
KEYWORDS
• bioenergy
• biofuels
• fossil fuels
• greenhouse gases
• production systems
• variety ROGER
REFERENCES
1. Rezk, H., Nassef, A. M., Inayat, A., Sayed, E. T., Shahbaz, M., & Olabi, A. G., (2019).
Improving the environmental impact of palm kernel shell through maximizing its
production of hydrogen and syngas using advanced artificial intelligence. Sci. Total
Environ., 658, 1150–1160.
2. Nazir, M. S., Mahdi, A. J., Bilal, M., Sohail, H. M., Ali, N., & Iqbal, H. M., (2019).
Environmental impact and pollution-related challenges of renewable wind energy
paradigm: A review. Sci. Total Environ., 683, 436–444.
3. Adler, P. R., Spatari, S., D’Ottone, F., Vazquez, D., Peterson, L., Del, G. S. J., Baethgen,
W. B., & Parton, W. J., (2018). Legacy effects of individual crops affect N2O emissions
accounting within crop rotations. GCB Bioenergy, 10, 123–136.
4. Holmberg, K., & Erdemir, A., (2019). The impact of tribology on energy use and CO2
Apple Academic Press
emission globally and in combustion engine and electric cars. Tribol. Int., 135, 389–396.
5. Leyva, A., & De Gauna, G. R., (2011). Cultivos energéticos y biocombustibles. Lychnos,
6, 44–49.
6. Toklu, E., (2017). Biomass energy potential and utilization in Turkey. Renewable Energy,
107, 235–244.
7. Zabed, H., Sahu, J. N., Suely, A., Boyce, A. N., & Faruq, G., (2017). Bioethanol
production from renewable sources: Current perspectives and technological progress.
Renew. Sustain. Energy Rev., 71, 475–501.
8. Ramachandra, T. V., & Hebbale, D., (2020). Bioethanol from macroalgae: Prospects and
Author Copy
challenges. Renewable and Sustainable Energy Reviews, 117, 109479.
9. Valdés-Rodríguez, O. A., & Palacios-Wassenaar, O. M., (2016). Evolution and current
situation of plantations for biofuel: perspectives and challenges for México. Agropro-
ductividad, 9, 33–41.
10. Dar, R. A., Dar, E. A., Kaur, A., & Phutela, U. G., (2018). Sweet sorghum-a promising
alternative feedstock for biofuel production. Renewable Sustainable Energy Rev., 82,
4070–4090.
11. Jiang, D., Hao, M., Fu, J., Liu, K., & Yan, X., (2019). Potential bioethanol production
from sweet sorghum on marginal land in China. J. Cleaner Prod., 220, 225–234.
12. Mengistu, G., Shimelis, H., Laing, M., Lule, D., & Mathew, I., (2020). Genetic variability
among Ethiopian sorghum landrace accessions for major agro-morphological traits and
anthracnose resistance. Euphytica, 216, 1–15.
13. Shukla, S., Felderhoff, T. J., Saballos, A., & Vermerris, W., (2017). The relationship
between plant height and sugar accumulation in the stems of sweet sorghum (Sorghum
bicolor (L.) Moench). Field Crop. Res., 203, 181–191.
14. Mathur, S., Umakanth, A. V., Tonapi, V. A., Sharma, R., & Sharma, M. K., (2017). Sweet
sorghum as biofuel feedstock: Recent advances and available resources. Biotechnol.
Biofuels, 10, 146.
15. Bojović, R., Popović, V. M., Ikanović, J., Živanović, L., Rakaščan, N., Popović, S.,
Ugrenović, V., & Simić, D., (2019). Morphological characterization of sweet sorghum
genotypes across environments. JAPS: J. Ani. Plant. Sci., 29, 721–729.
16. Ou, M. S., Awasthi, D., Nieves, I., Wang, L., Erickson, J., Vermerris, W., Ingram, L. O.,
& Shanmugam, K. T., (2016). Sweet sorghum juice and bagasse as feedstocks for the
production of optically pure lactic acid by native and engineered Bacillus coagulans
strains. BioEnergy Res., 9, 123–131.
17. Pabendon, M. B., Efendi, R., Santoso, S. B., & Prastowo, B., (2017). Varieties of sweet
sorghum super-1 and super-2 and its equipment for bioethanol in Indonesia. IOP Conference
Series: Earth and Environmental Science (Vol. 65). International conference on biomass:
technology, application, and sustainable development-2016; Bogor, Indonesia.
18. Adimassu, Z., Alemu, G., & Tamene, L., (2019). Effects of tillage and crop residue
management on runoff, soil loss and crop yield in the humid highlands of Ethiopia.
Agric. Syst., 168, 11–18.
19. Thivierge, M. N., Chantigny, M. H., Bélanger, G., Seguin, P., Bertrand, A., & Vanasse,
A., (2015). Response to nitrogen of sweet pearl millet and sweet sorghum grown for
ethanol in eastern Canada. Bioenergy Res., 8, 807–820.
20. Ding, N., Yang, Y., Cai, H., Liu, J., Ren, L., Yang, J., & Xie, G. H., (2017). Life cycle
assessment of fuel ethanol produced from soluble sugar in sweet sorghum stalks in
North China. J. Clean. Prod., 161, 335–344.
21. Phukoetphim, N., Chan-u-tit, P., Laopaiboon, P., & Laopaiboon, L., (2019). Improvement
Apple Academic Press
of bioethanol production from sweet sorghum juice under very high gravity fermentation:
Effect of nitrogen, osmoprotectant, and aeration. Energies, 12, 3620.
22. Garofalo, P., Campi, P., Vonella, A. V., & Mastrorilli, M., (2018). Application of multi-
metric analysis for the evaluation of energy performance and energy use efficiency of
sweet sorghum in the bioethanol supply-chain: A fuzzy-based expert system approach.
App. Energy, 220, 313–324.
23. Draghici, I., Draghici, R., Diaconu, A., Croitoru, M., Paraschiv, A. N., Dima, M.,
Ciuciuc, E., & Ciuciuc, D., (2019). Results on bioenergetic potential of some sweet
sorghum hybrids cultivated under psamosols conditions in Southern Oltenia. In: E3S
Author Copy
Web of Conferences (Vol. 112). EDP Sciences.
24. Martiwi, I. N. A., Nugroho, L. H., Daryono, B. S., & Susandarini, R., (2020).
Morphological variability and taxonomic relationship of Sorghum bicolor (L.) moench
accessions based on qualitative characters. Annu. Res. Rev. Biol., 35, 40–52.
25. Hilley, J. L., Weers, B. D., Truong, S. K., McCormick, R. F., Mattison, A. J., McKinley,
B. A., Morishige, D. T., & Mullet, J. E., (2017). Sorghum Dw2 encodes a protein kinase
regulator of stem internode length. Sci. Rep., 7, 1–13.
26. Lazarides, M., Hacker, J. B., & Andrew, M. H., (1991). Taxonomy, cytology and ecology
of indigenous Australian sorghums (Sorghum Moench: Andropogoneae: Poaceae). Aust.
Syst. Bot., 4, 591–635.
27. Saballos, A., (2008). Development and utilization of sorghum as a bioenergy crop. In:
Genetic Improvement of Bioenergy Crops (pp. 211–248). Springer, New York, NY,
(2008).
28. Mundia, C. W., Secchi, S., Akamani, K., & Wang, G., (2019). A regional comparison
of factors affecting global sorghum production: The case of North America, Asia and
Africa’s Sahel. Sustainability, 11, 2135.
29. FAO, (2020). Food and Agriculture Organization of the United Nations. http://www.
fao.org/faostat/es/#data/QC (accessed on 28 October 2021).
30. Appiah-Nkansah, N. B., Li, J., Rooney, W., & Wang, D., (2019). A review of sweet
sorghum as a viable renewable bioenergy crop and its techno-economic analysis.
Renewable Energy, 143, 1121–1132.
31. Rutto, L. K., Xu, Y., Brandt, M., Ren, S., & Kering, M. K., (2013). Juice, ethanol, and
grain yield potential of five sweet sorghum (Sorghum bicolor [L.] moench) cultivars. J.
Sustain. Bioenergy Syst., 3, 113–118.
32. Briand, C. H., Geleta, S. B., & Kratochvil, R. J., (2018). Sweet sorghum (Sorghum
bicolor [L.] Moench) a potential biofuel feedstock: Analysis of cultivar performance in
the Mid-Atlantic. Renewable Energy, 129, 328–333.
33. Perrin, R., Fulginiti, L. E., Bairagi, S., & Dweikat, I., (2018). Sweet Sorghum as feedstock
in great plains corn ethanol plants: The role of biofuel policy. J. Agr. Resour. Econ., 43,
34–45.
34. Anami, S. E., Zhang, L. M., Xia, Y., Zhang, Y. M., Liu, Z. Q., & Jing, H. C., (2015). Sweet
sorghum ideotypes: Genetic improvement of stress tolerance. Food Energy Secur., 4, 3–24.
35. Bonin, C. L., Heaton, E. A., Cogdill, T. J., & Moore, K. J., (2016). Management of sweet
sorghum for biomass production. Sugar Tech., 18, 150–159.
36. Larnaudie, V., Rochon, E., Ferrari, M. D., & Lareo, C., (2016). Energy evaluation of fuel
bioethanol production from sweet sorghum using very high gravity (VHG) conditions.
Renewable Energy, 88, 280–287.
37. Jardim, A. M. D. R. F., Da Silva, G. Í. N., Biesdorf, E. M., Pinheiro, A. G., Da Silva, M.
Apple Academic Press
V., Araújo, J., Dos, S. A., et al., (2020). Production potential of Sorghum bicolor (L.)
Moench crop in the Brazilian semiarid. PUBVET, 14, 1–13.
38. Umakanth, A. V., Kumar, A. A., Vermerris, W., & Tonapi, V. A., (2019). Sweet sorghum
for biofuel industry. In: Breeding Sorghum for Diverse End Uses (pp. 255–270).
Woodhead Publishing.
39. Maw, M. J., Houx, III. J. H., & Fritschi, F. B., (2016). Sweet sorghum ethanol yield
component response to nitrogen fertilization. Ind. Crops Prod., 84, 43–49.
40. Xuan, T. D., Phuong, N. T., Khang, D. T., & Khanh, T. D., (2015). Influence of sowing
times, densities, and soils to biomass and ethanol yield of sweet sorghum. Sustainability,
Author Copy
7, 11657–11678.
41. Zuo, W., Gu, C., Zhang, W., Xu, K., Wang, Y., Bai, Y., & Dai, Q., (2019). Sewage sludge
amendment improved soil properties and sweet sorghum yield and quality in a newly
reclaimed mudflat land. Sci. Total. Environ., 654, 541–549.
42. Ramírez-Jaramillo, G., (2020). Agroclimatic conditions for growing Sorghum bicolor L.
moench, under irrigation conditions in Mexico. Open Access Library Journal, 7, 1–14.
43. Aguilar-Sánchez, P., Navarro-Pineda, F. S., Sacramento-Rivero, J. C., & Barahona-
Pérez, L. F., (2018). Life-cycle assessment of bioethanol production from sweet sorghum
stalks cultivated in the state of Yucatan, Mexico. Clean. Technol. Environ. Policy, 20,
1685–1696.
44. Han, K. J., Alison, M. W., Pitman, W. D., Day, D. F., Kim, M., & Madsen, L., (2012).
Planting date and harvest maturity impact on biofuel feedstock productivity and
quality of sweet sorghum grown under temperate Louisiana conditions. Agron. J., 104,
1618–1624.
45. Houx, III. J. H., & Fritschi, F. B., (2015). Influence of late planting on light interception,
radiation use efficiency and biomass production of four sweet sorghum cultivars. Ind.
Crops Prod., 76, 62–68.
46. Wolabu, T. W., Zhang, F., Niu, L., Kalve, S., Bhatnagar-Mathur, P., Muszynski, M. G.,
& Tadege, M., (2016). Three flowering locus T-like genes function as potential florigens
and mediate photoperiod response in sorghum. New Phytol., 210, 946–959.
47. Almodares, A., & Hadi, M. R., (2009). Production of bioethanol from sweet sorghum: A
review. Afr. J. Agr. Res., 4, 772–780.
48. Uchimiya, M., Knoll, J. E., Anderson, W. F., & Harris-Shultz, K. R., (2017). Chemical
analysis of fermentable sugars and secondary products in 23 sweet sorghum cultivars. J.
Agric. Food Chem., 65, 7629–7637.
49. Tang, C., Sun, C., Du, F., Chen, F., Ameen, A., Fu, T., & Xie, G. H., (2018). Effect of
plant density on sweet and biomass Sorghum production on semiarid marginal land.
Sugar Tech., 20, 312–322.
50. Sahu, H., Tomar, G. S., & Thakur, V. S., (2020). Effects of planting density and levels of
nitrogen on growth and development of sweet sorghum [Sorghum bicolor (L.) Moench]
varieties. J. Pharmacogn. Phytochem., 9, 1894–1898.
51. Tian, F., Bradbury, P. J., Brown, P. J., Hung, H., Sun, Q., Flint-Garcia, S., Rocheford, T.
R., et al., (2011). Genome-wide association study of leaf architecture in the maize nested
association mapping population. Nature Genetics, 43, 159–162.
52. Zhang, N., Van, W. A., He, L., Evers, J. B., Anten, N. P., & Marcelis, L. F., (2020). Light
from below matters: Quantifying the consequences of responses to far-red light reflected
upwards for plant performance in heterogeneous canopies. Plant. Cell. Environ., 1–12.
53. Choi, Y., Han, H., Shin, S., Heo, B., Choi, K., & Kwon, S., (2019). Effect of planting
Apple Academic Press
Author Copy
(2018). Cover crop and nitrogen fertilization influence soil carbon and nitrogen under
bioenergy sweet sorghum. Agron. J., 110, 463–471.
57. Ameen, A., Yang, X., Chen, F., Tang, C., Du, F., Fahad, S., & Xie, G. H., (2017).
Biomass yield and nutrient uptake of energy sorghum in response to nitrogen fertilizer
rate on marginal land in a semi-arid region. BioEnerg. Res., 10, 363–376.
58. Kering, M. K., Temu, V. W., & Rutto, L. K., (2017). Nitrogen fertilizer and panicle
removal in Sweet Sorghum production: Effect on biomass, juice yield and soluble sugar
content. J. Sustain Bioenergy Syst., 7, 14–26.
59. Sowiński, J., Konieczny, M., & Jama-Rodzeńska, A., (2018). The effect of yararega
fertilization on the nitrogen effectiveness and yield of sweet sorghum (Sorghum bicolor
(L.) moench). Acta Sci. Pol. Agric., 16, 235–246.
60. Maw, M. J., Houx, III. J. H., & Fritschi, F. B., (2019). Nitrogen content and use efficiency
of sweet sorghum grown in the lower Midwest. Agron. J., 111, 2920–2928.
61. Esperbent, C. E., (2015). Malezas: El Desafío Para el Agro Que Viene. [Online], 12,
235–240 http://www.redalyc.org/pdf/864/86443147004.pdf (accessed on 28 October
2021).
62. Vázquez-Navarro, J. M., Carrillo-Aguilera, J. C., & Cisneros-Flores, B. A., (2016). A
population study in a forage sorghum crop infested with sugarcane aphid Melanaphis
sacchari (Zehntner, 1897) (Hemiptera: Aphididae) at the Comarca Lagunera Region.
Entomología Mexicana, 3, 395–400.
63. Tejeda-Reyes, M. A., Díaz-Nájera, J. F., Rodríguez-Maciel, J. C., Vargas-Hernández,
M., Solís-Aguilar, J. F., Ayvar-Serna, S., & Flores-Yáñez, J. A., (2017). Evaluación en
campo de insecticidas sobre melanaphis sacchari (Zehntner) 1 en sorgo. Southwest.
Entomol., 42, 545–551.
64. Bowling, R. D., Brewer, M. J., Kerns, D. L., Gordy, J., Seiter, N., Elliott, N. E., Butin,
G. D., et al., (2016). Sugarcane aphid (Hemiptera: Aphididae): A new pest on sorghum
in North America. J. Integr. Pest. Manage., 7, 12–25.
65. Rodríguez-del-Bosque, L. A., & Terán, A. P., (2015). Melanaphis sacchari (Hemiptera:
Aphididae): A new sorghum insect pest in Mexico. Southwest. Entomol., 40, 433–435.
66. Dercas, N., & Liakatas, A., (2018). Sweet Sorghum Canopy Development in Relation to
Radiation and Water Use. Environ. Process, 5, 413–425.
67. Mengistu, M. G., Steyn, J. M., Kunz, R. P., Doidge, I., Hlophe, H. B., Everson, C. S.,
Jewitt, G. P. W., & Clulow, A. D., (2016). A preliminary investigation of the water use
efficiency of sweet sorghum for biofuel in South Africa. Water SA, 42, 152–160.
68. Prasad, S., Sheetal, K. R., Renjith, P. S., Kumar, A., & Kumar, S., (2019). Sweet
sorghum: An excellent crop for renewable fuels production. In: Rastegari, A., Yadav, A.,
& Gupta, A., (eds.), Prospects of Renewable Bioprocessing in Future Energy Systems:
Biofuel Bior. Technol. (Vol. 10, pp. 291–314), Springer, Cham.
Apple Academic Press
69. Lopez, J. R., Erickson, J. E., Asseng, S., & Bobeda, E. L., (2017). Modification of the
CERES grain sorghum model to simulate optimum sweet sorghum rooting depth for
rainfed production on coarse textured soils in a sub-tropical environment. Agric. Water
Managet., 181, 47–55.
70. Costa, G. H. G., Ciaramello, S., Giachini, J. W., Gazzola, W. C. B., Giachini, L. E.,
& Uribe, R. A. M., (2018). Effects of sweet sorghum harvest systems on raw material
quality. Sugar Tech., 20, 730–733.
71. Vlachos, C. E., Pavli, O. I., Flemetakis, E., & Skaracis, G. N., (2020). Exploiting pre-and
post-harvest metabolism in sweet sorghum genotypes to promote sustainable bioenergy
Author Copy
production. Ind. Crops. Prod., 155, 112758.
72. Oyier, M. O., Owuoche, J. O., Oyoo, M. E., Cheruiyot, E., Mulianga, B., & Rono, J.,
(2017). Effect of harvesting stage on sweet sorghum (Sorghum bicolor L.) genotypes in
western Kenya. Sci. World. J., 2017, 1–10.
73. Dutra, E. D., Alencar, B. R. A., Galdino, J. J., Tabosa, J. N., Menezes, R. S. C., De Araújo,
F. R. N., Dário, C. P., et al., (2018). First and second generation of ethanol production for
five sweet sorghum cultivars during soft dough grain. J. Exp. Agric., 25, 1–12.
74. López-Sandin, I., Gutiérrez-Soto, G., Gutiérrez-Díez, A., Medina-Herrera, N., Gutiérrez-
Castorena, E., & Zavala-García, F., (2019). Evaluation of the use of energy in the
production of sweet sorghum (Sorghum Bicolor (L.) Moench) under different production
systems. Energies, 12, 1713.
75. Ebrahimiaqda, E., & Ogden, K. L., (2018). Evaluation and modeling of bioethanol yield
efficiency from sweet sorghum juice. BioEnergy Res., 11, 449–455.
76. Wu, X., Staggenborg, S., Propheter, J. L., Rooney, W. L., Yu, J., & Wang, D., (2010).
Features of sweet sorghum juice and their performance in ethanol fermentation. Ind.
Crops. Prod., 31, 164–170.
77. Kumar, C. G., Rao, P. S., Gupta, S., Malapaka, J., & Kamal, A., (2015). Chemical
preservatives-based storage studies and ethanol production from juice of sweet sorghum
cultivar, ICSV 93046. Sugar tech., 17, 404–411.
78. Bridgers, E. N., Chinn, M. S., Veal, M. W., & Stikeleather, L. F., (2011). Influence of juice
preparations on the fermentability of sweet sorghum. Biol. Engineer. Trans., 4, 57–67.
79. Wang, M., Chen, Y., Xia, X., Li, J., & Liu, J., (2014). Energy efficiency and environmental
performance of bioethanol production from sweet sorghum stem based on life cycle
analysis. Bioresour. Technol., 163, 74–81.
80. Jankowski, K. J., Sokólski, M. M., Dubis, B., Załuski, D., & Szempliński, W., (2020).
Sweet sorghum-Biomass production and energy balance at different levels of agricultural
inputs. A six-year field experiment in north-eastern Poland. Eur. J. Agron., 119, 126119.
81. Glab, L., Sowiński, J., Chmielewska, J., Prask, H., Fugol, M., & Szlachta, J., (2019).
Comparison of the energy efficiency of methane and ethanol production from sweet
sorghum (Sorghum bicolor (L.) Moench) with a variety of feedstock management
technologies. Biomass Bioenergy, 129, 105332.
82. Argote, F. E., Cuervo, R. A., Osorio, E., Ospina, J. D., & Castillo, H. S. V., (2015).
Evaluation of ethanol production from molasses with native strains of Saccharomyces
cerevisiae. Biotecnol. Sector Agropecuario Agroind, 13, 40–48.
83. Buruiană, C. T., Vizireanu, C., & Furdui, B., (2018). Bioethanol production from sweet
sorghum stalk juice by ethanol-tolerant Saccharomyces cerevisiae strains: An overview.
The Annals of the University Dunarea de Jos of Galati. Fascicle VI-Food Technology,
42, 153–167.
Apple Academic Press
84. Pilap, W., Thanonkeo, S., Klanrit, P., & Thanonkeo, P., (2018). The potential of the
newly isolated thermotolerant Kluyveromyces marxianus for high-temperature ethanol
production using sweet sorghum juice. 3 Biotech, 8, 126.
85. Laopaiboon, P., Khongsay, N., Phukoetphim, N., & Laopaiboon, L., (2019). Ethanol
production from sweet sorghum juice under very high gravity fermentation by Saccha-
romyces cerevisiae: Aeration strategy and its effect on yeast intracellular composition.
Chiang Mai J. Sci., 46, 481–494.
86. Vanderlip, R. L., & Reeves, H. E., (1972). Growth stages of sorghum [Sorghum bicolor,
(L.) moench.]. Agron. J., 64, 13–16.
Author Copy
87. De Las, C. H. R. M., Rodríguez, T. H., Paneque, P. R., & Díaz, M. A., (2011). Energy
cost of the knife roller CEMA 1400 for vegetable covering. Rev. Cienc. Téc. Agropecu.,
20, 53–56.
88. Paneque, P., & Rodríguez, Y. S., (2006). Energy cost of the rice mechanized harvest in
Cuba. Cienc. Téc. Agropecu., 15, 19–23.
89. Bridges, T. C., & Smith, E. M., (1979). A method for determining the total energy input
for agricultural practices. Trans. ASAE., 2, 781–0784.
90. Stout, B. A., (1990). Handbook of Energy for World Agriculture (1st edn.). Texas A & M
University: College Station, TX, USA, pp. 50–95, ISBN: 1-85166-349-5.
91. Crowell, E. A., & Ough, C. S., (1979). A modified procedure for alcohol determination
by dichromate oxidation. Am. J. Enol. Viticult., 30, 61–63.
92. Khalil, S. R., Abdelhafez, A. A., & Amer, E. A. M., (2015). Evaluation of bioethanol
production from juice and bagasse of some sweet sorghum varieties. Ann. Agric. Sci.,
60, 317-324.
93. Miller, G. L., (1959). Use of dinitrosalicylic acid reagent for determination of reducing
sugars. Anal. Chem., 31, 426–428.
94. Disasa, T., Feyissa, T., & Admassu, B., (2017). Characterization of Ethiopian sweet
sorghum accessions for 0 brix, morphological and grain yield traits. Sugar Tech., 19,
72–82.
95. Sarungallo, R. S., Melawaty, L., Djonny, M., Bulo, L., Mangera, L., Pabendon, M. B., &
Sarungallo, Z. L., (2019). Fermentation juice sweet sorghum genotip 4-183A using batch
system by optimizing the concentration of inoculum and substrate. The 1st international
conference on education and technology (ICETECH). IOP Conf. Series: Journal of
Physics: Conf. Series 1464 (2020), 012050.
96. Harlan, J. R., & de Wet, J. M. J. (1972). A simplified classification of cultivated sorghum.
Crop Science, 12, 172–176.
97. Bogunovic, I., Pereira, P., Kisic, I., Sajko, K., & Srakac, M. (2018). Tillage management
impacts on soil compaction, erosion and crop yield in Stagnosols (Croatia). CATENA
160, 376–384.
Biotechnology Development of
Bioethanol from Sweet Sorghum Bagasse
DANIEL TINÔCO
Biochemical Engineering Department, School of Chemistry,
Author Copy
Federal University of Rio de Janeiro, Rio de Janeiro–21941909, RJ, Brazil,
E-mail: dneto@peq.coppe.ufrj.br
ABSTRACT
Sorghum is one of the most cultivated cereals in the world, with great
potential for energy and fuel production. Its lignocellulosic fraction can
be used for cellulosic ethanol production, after undergoing treatment steps
to make sugars available. This chapter presents the most recent trends in
the bioprocess for the ethanol production from sweet sorghum bagasse,
highlighting the main physical, chemical, physical-chemical, and biological
pretreatments used for cellulose digestibility such as acid-base, SE, ammonia
fiber expansion (AFEX), organic solvents, ILs, microwave, and combined
methods. Saccharification and fermentation approaches were also presented,
such as simultaneous (SSF), separate (SHF), and co-fermentation (SScF)
processes. Finally, a biotechnological evolution of sweet sorghum bagasse
was presented from the main scientific reports in the last 12 years on its use
as a raw material for the fuel cellulosic ethanol production.
12.1 INTRODUCTION
Sorghum is one of the most cultivated cereals in the world, along with
wheat, rice, corn, and barley [1], being composed on average of 50–60%
lignocellulosic biomass (LCB) [2]. Considered a viable energy crop for
alcohol fuel production, sorghum, especially the sweet variety, has great
potential for the generation of cellulose-based products such as ethanol,
Author Copy
extraction, and decortication. Chemical methods are based on dilute acids
and alkalis, organic solvents, ILs, and oxidative agents. Physical-chemical
treatments combine mechanical and chemical processes, being the SE, liquid
hot water (LHW), supercritical CO2, AFEX, and microwave the most used
methods. Biological treatment can be combined with other methodologies,
being marked by the use of enzymes and microorganisms (bacteria and fungi)
capable of degrading the lignocellulosic material [5–7].
The next step is the enzymatic hydrolysis and fermentation. These
processes can be conducted in different approaches: simultaneously in the
same bioreactor or separately in different bioreactors [7]. SSF is generally
conducted at low temperatures due to the thermal tolerance of the microor-
ganisms used. However, this condition limits the fermentation efficiency,
requiring a pre-saccharification step under optimized conditions. Once the
sugars are released, the temperature can be readjusted to the ideal fermentation
condition and, then, the microorganism inoculation can be completed. This
process is classified as delayed simultaneous saccharification and fermenta-
tion (DSSF) [8]. The SHF has the advantage that the saccharification and
fermentation steps can be conducted under ideal individualized conditions.
Meanwhile, the ethanol yield and productivity, and the bioprocess economy
tend to be lower than those obtained with the SSF approach. Simultaneous
saccharification and co-fermentation (SScF) are similar to the SSF process,
with the additional fermentation of free sugars present in the sweet sorghum
stem. Therefore, the C5 and C6 sugars can be simultaneously converted to
ethanol [9].
This chapter aimed to present the main trends related to the cellulosic
ethanol biotechnology from sweet sorghum bagasse, highlighting the treatment
methodologies and the most used saccharification and fermentation approaches
in recent years, especially between 2009 and 2020. The biotechnological
evolution was also presented for the cellulosic ethanol production, highlighting
the most used combinations of pretreatment and sweet sorghum fermentation,
as well as the microorganisms and most relevant aspects involved in the sweet
sorghum ethanol bioprocess.
Apple Academic Press
12.2 SORGHUM
Sorghum is the fifth most cultivated cereal in the world, with an expected
Author Copy
production of 327 Mtons by 2027 [10]. The main producing regions are
Africa (40.1%) and the Americas (38.1%), where sorghum is used in human
food and for the alcoholic drinks and biofuels production. Next are Asia
(17.3%), Oceania (3.2%), and Europe (1.4%) (Figure 12.1).
world between 1998 and 2018, four are African, three American, two Asian,
and one from Oceania (Figure 12.2). The United States is the largest sorghum
producer in the world, with an average production of 10.6 Mtons, followed
by Nigeria with around 7.4 Mtons, India with 6.65 Mtons, and Mexico with
Apple Academic Press
6.1 Mtons. Sudan and Sudan (former) account for a production of 4.4 and 3.8
Mtons, respectively [1]. Argentina is a reference in South American sorghum
production with 3 Mtons, together with Brazil, which has widely cultivated
sorghum in association with sugarcane for bioethanol production, although
not among the 10 largest producers [11]. Since 1960, Ethiopia has increased
its sorghum production, with a recent amount of 2.9 Mtons. In addition to
India, China has stood out in Asia since the sorghum participation in its
Author Copy
domestic market increased. China has imported large cereal quantities in the
past, about 3 Mtons in 2012 and 18 Mtons in 2014 [10]. Today, China has
an average production of 2.6 Mtons, being the ninth largest producer in the
world and contributing, together with India, with more than 85% of Asian
sorghum production [12]. In Oceania, the only main contributor is Australia,
with a production of 1.9 Mtons.
The average world sorghum production and the corresponding planted
area for the period 1998–2018 was 63.1 Mtons and 45 Mha, respectively,
which represents a world average yield of 1.4 t/ha1 (Figure 12.3).
Author Copy
FIGURE 12.3 Area harvested and production of sorghum in the world [1].
Author Copy
tropical countries, sweet sorghum is widely cultivated in the rainy seasons
from June to July, and dry seasons from September to October, while in
temperate countries, the planting is limited to once a year [4]. In Brazil, where
sugarcane is the main feedstock for bioethanol production, sweet sorghum
has been cultivated in the sugarcane off-season, between November and
May, allowing the operational integration of the commercial distilleries [15].
In the USA, with their corn bioethanol, sweet sorghum has been planted
between May and June, which has contributed to biomass yields of 26 to 29
t/ha, with an ethanol production of at least 14,500 L/ha [16].
Author Copy
insoluble), fatty constituents (saturated and unsaturated), and vitamins [12].
Its calorific value (18.3 MJ/kg) is comparable with other lignocellulosic
feedstocks as switchgrass (18.4 MJ/kg) and big bluestem (18.6 MJ/kg) [4].
Due to its chemical and power characteristics, sweet sorghum bagasse can
be used for the paper and animal feed production, in soil applications as
a fertilizer, for the power co-generation, and as a raw material for fuel
cellulosic ethanol [17].
These costs normally increase with a decrease in the solids load of the treated
material (dilute solution form) [19].
The main pretreatments used with lignocellulosic materials include
physical, chemical, physical-chemical, and biological methods. Many of
Apple Academic Press
Author Copy
FIGURE 12.4 Main methodologies for the sweet sorghum bagasse treatment.
12.3.1.1 PHYSICAL
10 to 15% [21]. The grinding reduces the particle size by mechanical shear,
which can be classified according to their diameter by the sieving process.
The particle size is a parameter related to the material’s saccharification
efficiency, as it is inversely proportional to its surface area and, therefore,
represents the contact area available to the action of the hydrolysis agents. In
several studies, the size range of the sweet sorghum bagasse particles after
grinding and sieving was 0.2 to 1.2 mm. Cao et al. [22] used a three-roller
Author Copy
mill to separate broth and sweet sorghum bagasse, which was air-dried,
ground, and sieved at 0.42 mm. Chen et al. [23] obtained the sweet sorghum
bagasse by crushing in a roller press, drying at 20°C to maintain the final
moisture content at 20%, grinding, and sieving in three fractions: 9.5–18,
4–6, and 1–2 mm. Darkwah et al. [24] used a mechanical extractor to sepa-
rate the juice from the bagasse, which was air-dried and ground with a 1 mm
mesh in a knife mill. Lavudi et al. [25] submitted the sorghum bagasse to
oven drying to reduce the moisture content to 9–10%, being subsequently
chopped, ground, and sieved at 0.6 mm.
12.3.1.2 CHEMICAL
Author Copy
responsible for different yields of released sugar. The treatment with 0.5%
(v/v) H2SO4 at 180°C for 5 min allowed a release of 66% glucan, with total
xylan removal. Compared to untreated bagasse, the available glucan amount
increased by 51% with treatment based on diluted acid [28]. In another
investigation, an improvement in the enzymatic hydrolysis step was verified
with the application of 1.75% (v/v) H2SO4 at 121°C for 40 min, which
favored the release of 14.22 g/L xylose and 2.42 g/L glucose. A small amount
of inhibitory compounds was also produced: 1.34 g/L acetic acid, 0.90 g/L
phenol, 0.12 g/L hydroxymethylfurfural (HMF), and 0.98 g/L furfural [29].
In the treatment based on H3PO4 (85% v/v) at 40–85°C, it was not observed
the HMF and furfural formation. After acid pretreatment at 50°C for 43
min, followed by enzymatic hydrolysis, the released glucose yield was 85%
[30]. The diluted H2SO4 treatment was investigated at 150°C for 1 h, at high
pressure in four different concentrations: 0.5; 1.0, 1.5 and 2.0% (v/v). A
greater recovery of glucan and xylan, around 88% and 91%, respectively,
was observed in the pre-treatment with 2% (v/v) H2SO4 [31].
Basic treatment is based on the application of hydroxides (NaOH, KOH,
NH4OH and Ca (OH)2), Na2CO3 and similar alkaline compounds, usually at
140–200°C, for a residence time of minutes to hours [19], which are respon-
sible for the saponification reactions of the ester bonds between lignin and
other carbohydrates [18]. These reactions reduce the polymerization degree
and the lignocellulosic material crystallinity, due to the lignin destruction [27]
and the cellulose and hemicellulose swelling, with consequent degradation
of their crystals [12]. Basic treatment causes lesser sugar loss than acid treat-
ment, and it is usually carried out with diluted alkalis [27]. Different alkalis
have been used to treat sweet sorghum bagasse. Treatment with NaOH at
121°C, for 1 hour at 15 psi, led to an approximate 37% lignin removal, with
88% cellulose recovery. In this method, the furfural and HMF formation
was not verified [32]. The treatment based on alkaline distillation using 10%
(w/w) NaOH was carried out at 100°C, for 30 min at 0.04 MPa. A recovery
of approximately 94.5% glucan and 86% xylan, and a removal of about 72%
lignin were obtained [33]. The alkaline hydrogen-peroxide treatment was
Apple Academic Press
Author Copy
established by Yu et al. [33], but without biomass washing, increased the
fermentable sugar conversion from 44% to 65% [35]. The residual alkalis
use is an economical and environmental method alternative for the basic
treatment of sweet sorghum bagasse. The black liquor, resulting from the
basic pretreatment of the empty fruit cluster, was used at 150°C for 30 min.
This residue contained about 76% NaOH and pH 13, which contributed to
approximately 62% bagasse delignification [20]. The green liquor, generated
in kraft pulp mills, consisting of Na2CO3 and NaS2 (simulated composition),
was used at 160°C for 110 min, with 18% total titratable alkaline charge and
40% sulfidity. A sugar yield of 82.6% was achieved after this treatment [36].
The comminated treatment is an efficient alternative for the treatment
of the lignocellulosic material, especially between acid and base, since
together, these methodologies can remove lignin and hemicellulose, making
cellulose available with greater purity. The combination of 1.4% (v/v) H2SO4
at 120°C for 47 min, and 0.25 mol/L NaOH at 120C for 5 min, resulted in
92% hemicellulose removal 90% lignin removal, with only 2.3% cellulose
lost, whose composition in the bagasse after treatment was approximately
79% [37]. The acid treatment with 1.5% (v/v) H2SO4 in autoclave for 33
min followed by the basic treatment with 4% (v/v) H2O2 for 45 h, pH 11.5,
allowed almost 77% hemicellulose and lignin removal, making available
about 79% cellulose for the reducing sugars generation [38].
methanol, acetone, organic acids (formic and acetic), and ethylene glycol [19].
Although expensive, most organic solvents can be reused after recovery by
evaporation and other extractive technologies, thereby reducing the adverse
effects of their presence in the fermentation step [27]. When treated with 50%
Apple Academic Press
ethanol and 1% H2SO4, at 140°C for 30 min, sweet sorghum stem showed a
sugar yield of 77%, 2-fold greater than untreated bagasse [39]. Acetone treat-
ment at 180°C for 60 min, improved the enzymatic hydrolysis step by 94.2%
in a process intended for the joint production of acetone, butanol, and ethanol
(ABE process), allowing the removal of 143 g lignin and the release of 250 g
hemicellulose for each 1 Kg bagasse treated [40]. Concerning the biorefinery
concept, treatment with an aqueous solution of 60% (v/v) ethanol and 20%
Author Copy
(v/v) isopropanol at 140°C for 30 min, with 1% sulfuric acid as catalyst,
contributed to a hydrolysis yield of 90.3%, producing 24.9 g/L glucose [41].
Ionic liquids (ILs) are salts formed by a large volume organic cation and an
inorganic anion of different size, usually smaller than the cation, which are
liquids at room temperature [19]. ILs are considered special solvent types,
being classified as green solvents, as they do not release toxic or flammable
substances during the cellulose treatment. These compounds are capable of
dissolving lignin and other carbohydrates by forming hydrogen bonds with
the hydroxyls of the sugars present in the lignocellulosic material [42]. The
sweet sorghum bagasse treated with 1-Butyl-3-methylimiazolium chloride
([BMIM] Cl) at 110°C for 120 min, showed a reduction in the hemicellulose
composition from 26.3% in untreated bagasse to 16.7% after treatment
[43]. ILs such as 1-allyl-3-methylimdazolium acetate ([AMIM] OAc),
1-ethyl-3-methylimidazolium formate ([EMIM] Fmt) and 1-ethyl-3-methy-
limidazolium acetate ([EMIM] OAc) showed high capacity dissolving the
lignocellulosic fraction of sweet sorghum, without causing damage to the
cellulase enzymatic action. A sugar yield 7.5-fold higher was verified with
the treatment using [EMIM] Fmt, compared to untreated bagasse, which also
favored a bacterial nanocellulose productivity of 0.71 g/(L.d) [44].
12.3.1.3 PHYSICAL-CHEMICAL
Author Copy
pressures [45], often in the range of 160–270°C and 20–50 bar, respectively,
for a few seconds or minutes [12], without the addition of a chemical reagent.
Therefore, it is a lower environmental impact method, responsible for the
complete sugar recovery when compared to other pretreatments [27]. SE can
be associated with water and SO2 impregnations to increase the lignocellu-
losic fiber fracture efficiency. The application of the SE treatment increased
in 2.5-fold the maximum cellulose conversion and the glucose release from
sweet sorghum bagasse treated in a reactor with 0.25 t/h high pressure, at
160°C, for 5 min [43]. Sweet sorghum bagasse samples impregnated with
SO2 gas were treated with steam at 190C for 5 min. Approximately 54%
glucan, 10% xylan, and 26% lignin were obtained after treatment [46]. The
sweet sorghum bagasse submitted to steam at 200°C for 5 min released
leachate, which had its liquid and solid fractions separated by a cyclone. A
composition of 52% cellulose, 9% hemicellulose, and 25% lignin was veri-
fied for the released solid fraction [47]. Impregnation with water was also
investigated in the SE treatment of dry sweet sorghum bagasse. After being
subjected to steam at 215°C for 2 min, about 37.3% of total sugars could be
recovered from the biomass treated [48].
was treated at 200°C for 6.5 min, hemicellulose solubilization of 85% was
achieved, being greater than the 74% yield obtained with the treatment at
190°C for 20 min [49].
Apple Academic Press
Author Copy
reduction, hemicellulose depolymerization, and deacetylation of acetyl
groups [45]. In AFEX treatment, the fermentation inhibitors formation and
a liquid current are not verified being, therefore, considered a dry-to-dry
process [18]. Ammonia can be reused in this method, reducing process costs,
and avoiding environmental problems with its disposal [27]. The optimi-
zation of AFEX treatment conditions was defined as 140°C, 30 min, 2:1
for ammonia/biomass ratio, and a mixture content of 120%. Under these
conditions, the glucan and xylan conversion reached 90% for sweet sorghum
bagasse. Free sugars destruction was observed during the AFEX treatment,
which was solved with the previous biomass washing [50].
180 W for 20 min, equivalent to a power of 43.2 kJ/g of dry biomass, in the
sulfuric acid solution (50 g/Kg of bagasse) presence at 82°C, a yield of 820 g
sugar for each 1 Kg treated bagasse was obtained after the process [52]. The
microwave-assisted acid treatment at 180 W and microwave-assisted alkali at
Apple Academic Press
Author Copy
12.3.1.4 BIOLOGICAL
produced about 2.5-fold more fermentable sugars than the control assays
using untreated sweet sorghum bagasse in a meshless bioreactor and without
a humid airflow [57].
Apple Academic Press
Author Copy
(SHF). When the SSF process steps take place at relatively large intervals,
the process is classified as DSSF. Treated biomass can also be saccharified
and fermented together with sweet sorghum juice in a process classified as
SSCF (Figure 12.5).
FIGURE 12.5 Saccharifications and fermentation approaches for the cellulosic ethanol
production from sweet sorghum bagasse. [Abbreviations: SHF: separate hydrolysis and fermen-
tation; SSF: simultaneous saccharification and fermentation; DSSF: delayed simultaneous
saccharification and fermentation; SScF: simultaneous saccharification and co-fermentation].
the SSF process is limited by the recovery difficulty and reuse of the micro-
organisms and enzymes, also requiring distinct favorable conditions for
each step such as optimal temperature and pH [58]. Treated sweet sorghum
bagasse can be submitted to the SSF process, using different enzymatic
Apple Academic Press
loads and microorganisms, for the cellulosic ethanol production. The sweet
sorghum bagasse treated with diluted H2SO4 was fermented in two steps.
Firstly, Neurospora crassa DSM 1129 fungus was used to produce lignocel-
lulolytic enzymes. Posteriorly, the SSF process was performed using these
enzymes, with cellulase and β-glucosidase supplementation at 6 FPU/g.
About 27.6 g/L ethanol was produced, with a yield of 84.7% [59]. The
SSF process optimization was performed to the efficient cellulosic ethanol
Author Copy
production. At 37°C, 25 FPU/g enzymatic load, and 1.4 g/L Saccharomyces
cerevisiae ATCC 24858, about 38 g/L ethanol, with 89.4% yield and 1.28
g/L/h productivity, were produced from biomass treated with dilute H2SO4
[28]. Approximately 85 g/L ethanol was achieved in 21 h fermentation by
S. cerevisiae JP1 at 37°C, after 6 h enzymatic hydrolysis at 50°C, with 32.8
FPU/g enzymatic load, and combined acid-base pretreatment. An ethanol
yield of 63% and 4 g/L/h productivity were obtained [37]. At severe alkaline
treatment using NaOH and H2O2, about 19.3 g/L ethanol was obtained by
S. cerevisiae CICC1308 at 36°C, after the pre-saccharification at 60 FPU/g
cellulase and 80 IU/g β-glucosidase, at 50°C for 12 h. Ethanol yield reached
more than 88% in 48 h of SSF [34]. The combined acid-base treatment
followed by hydrolysis at 20 IU/g, at 60°C for 58 h, and fermentation at
35°C by Pichia kudriavzevii HOP-1 resulted in 26.8 g/L ethanol in 48 h. P.
kudriavzevii HOP-1 was considered advantageous for the SSF process due
to its thermotolerant capacity, which confers economic advantages to the
commercial cellulosic ethanol production [25]. The black liquor application
for 30 min followed by enzymatic hydrolysis with 30 FPU/g enzymatic load
and fermentation at 32°C by commercial S. cerevisiae resulted in 45.06 g/L
ethanol in 72 h [20]. The alkaline treatment based on Na2CO3, resulting from
the reaction between NaOH from the bagasse basic treatment and the CO2
released from the sweet sorghum juice fermentation, was associated with
saccharification and fermentation at 32°C for 72 h. A theoretical ethanol
yield of almost 82% glucan by S. cerevisiae was obtained [60]. The DSSF
process was also investigated after hydrothermal treatment associated with
liquefaction-saccharification at 50°C, with 10 FPU/g enzymatic load, for 24
h, and addition of extra enzymes. About 48.3 g/L ethanol by baker’s yeast at
30°C was reached. The ethanol yield and productivity under these conditions
were equal to 71.2% and 2.2 g/L/h, respectively [61].
equal to 25.7 g/L. Fed-batch SSF improved the final ethanol production and
contributed to reduce the cell inhibition caused by the high fermentation
medium viscosity containing a high solid load content [24].
SSF process can also be combined with biological pretreatment by solid-
state fermentation of sweet sorghum bagasse, in which filamentous fungi are
used to degrade the lignocellulosic fraction. In a study using Mucor indicus
CCUG 22424, the solid-state fermentation product was submitted to the SSF
Author Copy
process, with 15 FPU/g cellulase and 30 IU/g β-glucosidase, and fermen-
tation at 37°C for 48 h. Compared to the SSF process without biological
pretreatment, the solid-state fermentation allowed a 4.3-fold higher ethanol
yield, equal to 85.6% [62].
The separate hydrolysis and fermentation (SHF) process is widely used in the
ethanol production from sweet sorghum bagasse, although its yield achieved
is slightly lower than the SSF process yield [58]. The separation of these
steps allows each process to be conducted under optimal temperature and
pH conditions, thus contributing to an efficient conversion of cellulose and a
high ethanol production [8]. SHF process also allows the cellular biomass and
hydrolysate recovery, being widely used with filamentous fungi. In a study
using M. hiemalis CCUG 16148, the sweet sorghum bagasse treated by the
ultrasound-assisted NaOH was submitted to the SHF process at 32°C for 24 h,
and 81% yield and 0.70 g/L/h ethanol productivity have been achieved [63].
In another investigation, the sweet sorghum bagasse treated by microwave-
assisted diluted ammonia was submitted to 60 FPU/g cellulase and 64 CBU/g
hemicellulase, at 55°C for 24 h. After 48 h fermentation by S. cerevisiae
(D5A) at 30°C, about 21 g ethanol/100 g treated biomass were produced [23].
Cellulosic ethanol was also obtained by Dekkera bruxellensis GDB 248, yeast
capable of assimilating cellobiose and glucose. After the alkaline treatment
using H2O2 and enzymatic hydrolysis at 50°C, with 20 FPU/g commercial
cellulase, for 48 h, and ethanol yield of 0.44 g/g was achieved in 7 h fermenta-
tion at 32°C [64]. The acid treatment based on H2SO4 was optimized and used
for the sweet sorghum bagasse hydrolysis, whose hemicellulosic fraction
generated was submitted to fermentation at 30°C by Scheffersomyces stipitis
ATCC 58376. About 22 g/L ethanol, 0.40 g/g yield, and 0.34 g/L/h produc-
tivity were obtained after 55 h [29]. The cellulosic ethanol production was
also investigated without pre-treatment. For an enzymatic hydrolysis with
8.32 FPU/g load and a fermentation at 30°C by Trichosporon fermentans CBS
Apple Academic Press
439.83, approximately 20.7 g/L ethanol were obtained. The external nitrogen
supplementation increased ethanol production to 23.5 g/L, with 0.118 g/g
yield and 0.196 g/L/h productivity in 120 h [65].
Enzymatic hydrolysis step is usually performed in batch approaches.
However, some studies suggest that the fed-batch approaches can improve
the glucose release during the saccharification process. The sweet sorghum
bagasse treated with hot liquid water at 180°C, for 20 min at 4 MPa, was
Author Copy
submitted to saccharification in fed-batch, with supplemented Tween80.
The solid loads were fed in 24 h and 48 h, together with 20 FPU/g and 30
FPU/g cellulase at 50°C, to reach a final load of 20% and 30%, respectively.
Approximately 89 g/L glucose was released after 120 h, and, then, submitted
to fermentation by S. cerevisiae Y2034 at 30°C for 72 h. About 43.4 g/L
ethanol was produced [66].
SHF processes are generally compared to SSF processes, considering the
same biomass pretreatment, to identify and differentiate their advantages
and limitations. The SSF and SHF processes were used after sweet sorghum
bagasse steam-treated to produce cellulosic ethanol by S. cerevisiae (Tembec
1). In the SSF process, the hydrolysis step was conducted at 50°C for 12 h,
and the fermentation step at 37°C for 120 h. About 23.3 g/L ethanol was
produced, with a yield of almost 64%. In the SHF process, the hydrolysis step
was conducted at 50°C and the fermentation step at 30°C. Approximately
21.2 g/L ethanol was obtained, with a yield of almost 58% [46]. The ethanol
produced by P. kudriavzevii HOP-1 was investigated in the SSF and SHF
processes, after the acid-base pretreatment of sweet sorghum bagasse. Both
methods resulted in just over 26 g/L ethanol, however, the SSF productivity
(0.56 g/L/h) was 40% higher than the SHF productivity (0.40 g/L/h), due to
the time required for the maximum ethanol production, which was 48 h for
SSF and 66 h for SHF [25]. S. cerevisiae TISTR 5606 was used to convert
sweet sorghum bagasse treated by H2O2 and NaOH into cellulosic ethanol.
The saccharification and fermentation of the hydrolysate were carried out by
SSF, SHF, and DSSF. In the SSF process, the substrate, inoculum, and cellu-
lase were mixed in the same bioreactor, at 37°C, for 72 h, which resulted in
22.3 g/L ethanol. In the DSSF process, the hydrolysis step was carried out
at 50°C for 6 h, followed by fermentation at 37°C for 66 h. About 28.3 g/L
ethanol was produced by DSSF. The DSSF productivity was 26% higher
than the SSF and SHF productivities of 0.31 g/L/h [8].
Author Copy
for 24 h, which resulted in 94% conversion efficiency and 480 g ethanol/Kg
treated biomass. Concerning only lignocellulosic fraction, about 33 m³/ha
ethanol can be produced from sweet sorghum [52]. Z. mobilis ATCC 31821
and commercial S. cerevisiae were also used for the sorghum hydrolysate
fermentation, treated by microwave-assisted acid/base. For 10 g/L Z. mobilis
and 5 g/L S. cerevisiae, a yield of 0.50 g/g ethanol was obtained after 24 h
fermentation at 32°C [53].
All sweet sorghum components (grain, juice, and bagasse) can be used to
increase ethanol production from its lignocellulosic fraction, since the low
concentration of fermentative products is the main limitation to the industrial
production of the cellulosic ethanol [49]. The bagasse saccharification and
fermentation together with the juice fermentation can improve ethanol
yield from different feedstocks, making the bioprocess economically
advantageous. A previously treated and dehydrated sweet sorghum juice and
bagasse was responsible for 53 g/L ethanol production in 168 h. The nutrients
present in the sweet sorghum juice were beneficial to the fermentation
process, improving in 92% the final ethanol, in lesser time of 72 h [49]. The
fermentative integration was also performed for the sweet sorghum bagasse
obtained from the solid-state process. This biomass was treated by NaOH
and submitted to SScF by Z. mobilis TSH-ZM-01 at 32°C. Under optimized
conditions, about 69.5% yield was achieved, which can reduce the capital
costs and energy consumption, making the ethanol bioprocess commercially
viable on a large scale [33]. The bagasse, treated by SE-assisted dilute
Author Copy
wild-strain S. cerevisiae. The detoxification step was not necessary, which
reduces the production costs [69]. The co-fermentation of juice and bagasse
from sweet sorghum was also investigated by the cellular consortium
fermentation. Firstly, the lignocellulosic fraction was treated by H2SO4
(98% v/v) at 120°C for 1 h. Next, the treated biomass and the juice were
fermented at 30°C for 96 h, using S. cerevisiae ATCC 7754 and Z. mobilis
ATCC 29191. For each 1 Kg treated sweet sorghum (variety SS-301), about
160 mL ethanol was produced [14].
Cellulosic ethanol production from residual sweet sorghum has been exten-
sively studied in recent years, as a promising alternative to sugarcane ethanol.
In 2009–2020, several investigations were carried out, using different
treatment technologies and saccharification and fermentation approaches.
Approximately 32% of scientific articles investigated physical, chemical,
physical-chemical, and biological treatments to make cellulose digestibility
more efficient and to available the sweet sorghum bagasse sugars for fermen-
tation. Of the 68% scientific articles addressing the fermentation stage, about
54.3% corresponded to the SHF process, 30.4% to the SSF process, and
15.2% to the SScF process (Figure 12.6). Although lesser efficient than the
SSF process, the SHF process has been the most used method in recent scien-
tific research for converting sweet sorghum bagasse into ethanol. Possibly,
the conducting of saccharification and fermentation steps under respective
optimal conditions can explain this predominance. In turn, the promising
SScF process was the least reported method for sweet sorghum processing
since it is still a recent technology, starting its investigations in 2017.
Author Copy
FIGURE 12.6 Main saccharifications and fermentation approaches reported by scientific
articles in 2009–2020.
combined treatments used in the period. In 2017–2020, the acid, the alka-
line, and the acid-base treatments were also wildly investigated [8, 25, 29]
(Figure 12.8).
Author Copy
FIGURE 12.7 Main methods for the sweet sorghum treatment using the SSF approach.
FIGURE 12.8 Main methods for the sweet sorghum treatment using the SHF approach.
FIGURE 12.9
Author Copy
Main methods for the sweet sorghum treatment using the SScF approach.
Author Copy
FIGURE 12.10 Main producer microorganism of cellulosic ethanol from sweet sorghum
bagasse.
12.5 CONCLUSIONS
The bioprocess for the sweet sorghum bagasse ethanol production has
changed in recent years, especially due to the development of more effi-
cient, cheaper, and eco-friendlier treatment technologies and fermentation
approaches. The combination of different treatments will be increasingly
used, since it can take advantage of the separated methodologies benefits
to improve the ethanol conversion and yield. Furthermore, the simultaneous
saccharifications and fermentation processes using cellular consortia and the
free sugar co-fermentation can contribute to an optimized and economically
Author Copy
FIGURE 12.11 Evolution of the main methods for the sweet sorghum treatment.
KEYWORDS
• biotechnology evolution
• cellulosic ethanol
• fermentation approaches
• pretreatment methodologies
• saccharification and fermentation
• sweet sorghum bagasse
REFERENCES
1. Food and Agriculture Organization of the United Nations. (2018). FAOSTAT Crops.
http://www.fao.org/faostat/en/#data/QC/visualize (accessed on 28 October 2021).
2. Cutz, L., & Santana, D., (2014). Techno-economic analysis of integrating sweet sorghum
into sugar mills: The central American case. Biomass and Bioenergy, 68, 195–214. https://
doi.org/10.1016/j.biombioe.2014.06.011.
3. Yu, J., Zhang, T., Zhong, J., Zhang, X., & Tan, T., (2012). Biorefinery of sweet
sorghum stem. Biotechnol. Adv., 30(4), 811–816. https://doi.org/10.1016/j.biotechadv.
2012.01.014.
4. Appiah-Nkansah, N. B., Li, J., Rooney, W., & Wang, D., (2019). A review of sweet
Apple Academic Press
sorghum as a viable renewable bioenergy crop and its techno-economic analysis. Renew.
Energy, 143, 1121–1132. https://doi.org/10.1016/j.renene.2019.05.066.
5. Akhtar, N., Gupta, K., Goyal, D., & Goyal, A., (2016). Recent Advances in Pretreatment
Technologies for Efficient Hydrolysis of Lignocellulosic Biomass, 35(2), 489–511.
https://doi.org/10.1002/ep.
6. Lima, I., Bigner, R., & Wright, M., (2017). Conversion of sweet sorghum bagasse into
value-added biochar. Sugar Tech., 19(5), 553–561. https://doi.org/10.1007/s12355-
017-0508-8.
7. Stamenković, O. S., Siliveru, K., Veljković, V. B., Banković-Ilić, I. B., Tasić, M. B.,
Author Copy
Ciampitti, I. A., Đalović, I. G., et al., (2020). Production of biofuels from sorghum.
Renew. Sustain. Energy Rev., 124. https://doi.org/10.1016/j.rser.2020.109769.
8. Thanapimmetha, A., Saisriyoot, M., Khomlaem, C., Chisti, Y., & Srinophakun, P.,
(2019). A comparison of methods of ethanol production from sweet sorghum bagasse.
Biochem. Eng. J., 151, 107352. https://doi.org/10.1016/j.bej.2019.107352.
9. Chandel, A. K., Garlapati, V. K., Singh, A. K., Antunes, F. A. F., & Da Silva, S. S.,
(2018). The path forward for lignocellulose biorefineries: Bottlenecks, solutions, and
perspective on commercialization. Bioresour. Technol., 264, 370–381. https://doi.
org/10.1016/j.biortech.2018.06.004.
10. OECD-FAO, (2018). Chapter 3. Cereals Agricultural Outlook 2018–2027 (pp. 109–126).
OCED/FAO.
11. Mejila, D., & Lewis, B., (1999). Sorghum: Post-harvest operations. INPhO - Post-harvest
Compendium, 33.
12. Velmurugan, B., Narra, M., Rudakiya, D. M., & Madamwar, D., (2020). Sweet sorghum:
A potential resource for bioenergy production. In: Refining Biomass Residues for
Sustainable Energy and Bioproducts (pp. 215–242). Elsevier. https://doi.org/10.1016/
B978-0-12-818996-2.00010-7.
13. Cifuentes, R., Bressani, R., & Rolz, C., (2014). Energy for sustainable development the
potential of sweet sorghum as a source of ethanol and protein. Energy Sustain. Dev., 21,
13–19. https://doi.org/10.1016/j.esd.2014.04.002.
14. Khalil, S. R. A., Abdelhafez, A. A., & Amer, E. A. M., (2015). Evaluation of bioethanol
production from juice and bagasse of some sweet sorghum varieties. Ann. Agric. Sci.,
60(2), 317–324. https://doi.org/10.1016/j.aoas.2015.10.005.
15. Khawaja, C., Janssen, R., Rutz, D., Luquet, D., Trouche, G., Reddy, B., Rao, P.S.,
Basavaraj, G., Schaffert, R., Damasceno, C., et al. (2014). A Handbook of Energy Sorghum:
An alternative energy crop (WIP–Renewable Energies Sylvensteintr). Available at: http://
oar.icrisat.org/9049/1/Sweetfuel%20Handbook%20English%20version.pdf (accessed on
23 July 2020).
16. Bonin, C. L., Heaton, E. A., Cogdill, T. J., & Moore, K. J., (2016). Management of sweet
sorghum for biomass production. Sugar Tech, 18(2), 150–159. https://doi.org/10.1007/
s12355-015-0377-y.
17. Regassa, T. H., & Wortmann, C. S., (2014). Sweet sorghum as a bioenergy crop: Literature
review. Biomass and Bioenergy, 64, 348–355. https://doi.org/10.1016/j.biombioe.2014.
03.052.
18. Kumar, B., Bhardwaj, N., Agrawal, K., Chaturvedi, V., & Verma, P., (2020). Current
Perspective on Pretreatment Technologies Using Lignocellulosic Biomass: An Emerging
Biorefinery Concept, 199. https://doi.org/10.1016/j.fuproc.2019.106244.
19. Galbe, M., & Wallberg, O., (2019). Pretreatment for biorefineries: A review of common
Apple Academic Press
Author Copy
22. Cao, W., Sun, C., Liu, R., Yin, R., & Wu, X., (2012). Comparison of the effects of five
pretreatment methods on enhancing the enzymatic digestibility and ethanol production
from sweet sorghum bagasse. Bioresour. Technol., 111, 215–221. https://doi.org/10.1016/j.
biortech.2012.02.034.
23. Chen, C., Boldor, D., Aita, G., & Walker, M., (2012). Ethanol production from sorghum
by a microwave-assisted dilute ammonia pretreatment. Bioresour. Technol., 110,
190–197. https://doi.org/10.1016/j.biortech.2012.01.021.
24. Darkwah, K., Wang, L., & Shahbazi, A., (2016). Simultaneous saccharification
and fermentation of sweet sorghum after acid pretreatment. Energy Sources, Part A
Recover. Util. Environ. Eff., 38(10), 1485–1492. https://doi.org/10.1080/15567036.2
012.724146.
25. Lavudi, S., Oberoi, H. S., & Mangamoori, L. N., (2017). Ethanol production from sweet
sorghum bagasse through process optimization using response surface methodology. 3
Biotech, 7(4), 233. https://doi.org/10.1007/s13205-017-0863-x.
26. Rabemanolontsoa, H., & Saka, S., (2016). Various pretreatments of lignocellulosics.
Bioresour. Technol., 199, 83–91. https://doi.org/10.1016/j.biortech.2015.08.029.
27. Chen, H., Liu, J., Chang, X., Chen, D., Xue, Y., Liu, P., Lin, H., & Han, S., (2017). A
Review on the pretreatment of lignocellulose for high-value chemicals. Fuel Process.
Technol., 160, 196–206. https://doi.org/10.1016/j.fuproc.2016.12.007.
28. Wang, L., Luo, Z., & Shahbazi, A., (2013). Optimization of simultaneous saccharification
and fermentation for the production of ethanol from sweet sorghum (Sorghum bicolor)
bagasse using response surface methodology. Ind. Crop. Prod., 42, 280–291. https://doi.
org/10.1016/j.indcrop.2012.06.005.
29. Camargo, D., Sydney, E. B., Leonel, L. V., Pintro, T. C., & Sene, L., (2019). Dilute
acid hydrolysis of sweet sorghum bagasse and fermentability of the hemicellulosic
hydrolysate. Brazilian J. Chem. Eng., 36(1), 143–156. https://doi.org/10.1590/0104-
6632.20190361s20170643.
30. Lo, E., Brabo-Catala, L., Dogaris, I., Ammar, E. M., & Philippidis, G. P., (2020).
Biochemical conversion of sweet sorghum bagasse to succinic acid. J. Biosci. Bioeng.,
129(1), 104–109. https://doi.org/10.1016/j.jbiosc.2019.07.003.
31. Zhang, C., Wen, H., Chen, C., Cai, D., Fu, C., Li, P., Qin, P., & Tan, T., (2019).
Simultaneous saccharification and juice co-fermentation for high-titer ethanol production
using sweet sorghum stalk. Renew. Energy, 134, 44–53. https://doi.org/10.1016/j.renene.
2018.11.005.
32. Kim, M., Han, K., Jeong, Y., & Day, D. F., (2012). Utilization of Whole Sweet Sorghum
Containing Juice, Leaves, and Bagasse for Bio-Ethanol Production, 21(4), 1075–1080.
https://doi.org/10.1007/s10068-012-0139-5.
33. Yu, M., Li, J., Li, S., Du, R., Jiang, Y., Fan, G., Zhao, G., & Chang, S., (2014). A
Apple Academic Press
Author Copy
36. Pham, H. T. T., Nghiem, N. P., & Kim, T. H., (2018). Near theoretical saccharification
of sweet sorghum bagasse using simulated green liquor pretreatment and enzymatic
hydrolysis. Energy, 157, 894–903. https://doi.org/10.1016/j.energy.2018.06.005.
37. Barcelos, C. A., Maeda, R. N., Santa, A. L. M. M., & Pereira, N., (2016). Sweet sorghum
as a whole-crop feedstock for ethanol production. Biomass and Bioenergy, 94, 46–56.
https://doi.org/10.1016/j.biombioe.2016.08.012.
38. Guarneros-flores, J., Aguilar-uscanga, M. G., Morales-martínez, J. L., & López-zamora,
L., (2019). Maximization of Fermentable Sugar Production from Sweet Sorghum
Bagasse (Dry and Wet Bases) Using Response Surface Methodology (RSM), 633–639.
39. Ostovareh, S., Karimi, K., & Zamani, A., (2015). Efficient conversion of sweet sorghum
stalks to biogas and ethanol using organosolv pretreatment. Ind. Crop. Prod., 66,
170–177. https://doi.org/10.1016/j.indcrop.2014.12.023.
40. Jafari, Y., Amiri, H., & Karimi, K., (2016). Acetone pretreatment for improvement of
acetone, butanol, and ethanol production from sweet sorghum bagasse. Appl. Energy,
168, 216–225. https://doi.org/10.1016/j.apenergy.2016.01.090.
41. Nozari, B., Mirmohamadsadeghi, S., & Karimi, K. (2018). Bioenergy production from
sweet sorghum stalks via a biorefinery perspective. Appl. Microbiol. Biotechnol. 102(7),
3425–3438. https://doi.org/10.1007/s00253–018–8833–8.
42. Chen, D., Gao, D., Capareda, S. C., E, S., Jia, F., & Wang, Y., (2020). Influences of
hydrochloric acid washing on the thermal decomposition behavior and thermodynamic
parameters of sweet sorghum stalk. Renew. Energy, 148, 1244–1255. https://doi.org/
10.1016/j.renene.2019.10.064.
43. Zhang, J., Ma, X., Yu, J., Zhang, X., & Tan, T., (2011). The effects of four different
pretreatments on enzymatic hydrolysis of sweet sorghum bagasse. Bioresour. Technol.,
102(6), 4585–4589. https://doi.org/10.1016/j.biortech.2010.12.093.
44. Chen, G., Chen, L., Wang, W., & Hong, F. F., (2018). Evaluation of six ionic liquids
and application in pretreatment of sweet sorghum bagasse for bacterial nanocellulose
production. J. Chem. Technol. Biotechnol., 93(12), 3452–3461. https://doi.org/10.1002/
jctb.5703.
45. Batista Meneses, D., Montes de Oca-Vásquez, G., Vega-Baudrit, J. R., Rojas-Álvarez,
M., Corrales-Castillo, J., & Murillo-Araya, L. C. (2020). Pretreatment methods of
lignocellulosic wastes into value-added products: recent advances and possibilities.
Biomass Conv. Bioref. https://doi.org/10.1007/s13399–020–00722–0
46. Shen, F., Hu, J., Zhong, Y., Liu, M. L. Y., & Saddler, J. N., (2012). Ethanol production
from steam-pretreated sweet sorghum bagasse with high substrate consistency enzymatic
hydrolysis. JBB, 41, 157–164. https://doi.org/10.1016/j.biombioe.2012.02.022.
47. Pengilly, C., Diedericks, D., Brienzo, M., & Görgens, J. F., (2015). Enzymatic hydrolysis
Apple Academic Press
Author Copy
102, 211–219. https://doi.org/10.1016/j.apenergy.2012.03.039.
50. Li, B., Balan, V., Yuan, Y., & Dale, B. E., (2010). Process optimization to convert forage
and sweet sorghum bagasse to ethanol based on ammonia fiber expansion (AFEX)
pretreatment. Bioresour. Technol., 101(4), 1285–1292. https://doi.org/10.1016/j.biortech.
2009.09.044.
51. Choudhary, R., Loku, A., Liang, Y., & Siddaramu, T., (2012). Microwave pretreatment
for enzymatic saccharification of sweet sorghum bagasse. Biomass and Bioenergy, 39,
218–226. https://doi.org/10.1016/j.biombioe.2012.01.006.
52. Marx, S., Ndaba, B., Chiyanzu, I., & Schabort, C., (2014). Fuel ethanol production
from sweet sorghum bagasse using microwave irradiation. Biomass and Bioenergy, 65,
145–150. https://doi.org/10.1016/j.biombioe.2013.11.019.
53. Ndaba, B., Chiyanzu, I., Marx, S., & Obiero, G., (2014). Effect of saccharomyces
cerevisiae and Zymomonas mobilis on the co-fermentation of sweet sorghum bagasse
hydrolysates pretreated under varying conditions. Biomass and Bioenergy, 71, 350–356.
https://doi.org/10.1016/j.biombioe.2014.09.022.
54. Xu, Q., Zhao, M., Yu, Z., Yin, J., Li, G., Zhen, M., & Zhang, Q., (2017). Enhancing
enzymatic hydrolysis of corn cob, corn stover and sorghum stalk by dilute aqueous
ammonia combined with ultrasonic pretreatment. Ind. Crops Prod., 109, 220–226.
https://doi.org/10.1016/j.indcrop.2017.08.038.
55. Mishra, V., Jana, A. K., Jana, M. M., & Gupta, A., (2017). Improvement of selective
lignin degradation in fungal pretreatment of sweet sorghum bagasse using synergistic
CuSO4-syringic acid supplements. J. Environ. Manage., 193, 558–566. https://doi.org/
10.1016/j.jenvman.2017.02.057.
56. Mishra, V., Jana, A. K., Jana, M. M., & Gupta, A., (2017). Synergistic Effect of Syringic
Acid and Gallic Acid Supplements in Fungal Pretreatment of Sweet Sorghum Bagasse
for Improved Lignin Degradation and Enzymatic Saccharification. Process Biochem.
https://doi.org/10.1016/j.procbio.2017.02.011.
57. Mishra, V., & Jana, A. K., (2019). Sweet sorghum bagasse pretreatment by coriolus
versicolor in mesh tray bioreactor for selective delignification and improved sacchari-
fication. Waste and Biomass Valorization, 10(9), 2689–2702. https://doi.org/ 10.1007/
s12649-018-0276-z.
58. Olofsson, K., Bertilsson, M., & Lidén, G., (2008). A short review on SSF – an interesting
process option for ethanol production from lignocellulosic feedstocks. Biotechnol.
Biofuels, 1(1), 7. https://doi.org/10.1186/1754-6834-1-7.
59. Dogaris, I., Gkounta, O., & Mamma, D., (2012). Bioconversion of Dilute-Acid Pretreated
Sorghum Bagasse to Ethanol by Neurospora Crassa, 541–550. https://doi.org/10.1007/
s00253-012-4113-1.
60. Nghiem, N. P., & Toht, M. J., (2019). Pretreatment of sweet sorghum bagasse for ethanol
Apple Academic Press
Author Copy
580–585. https://doi.org/10.1016/j.indcrop.2013.06.024.
63. Goshadrou, A., Karimi, K., & Taherzadeh, M. J., (2011). Bioethanol production from
sweet sorghum bagasse by mucor hiemalis. Ind. Crops Prod., 34(1), 1219–1225. https://
doi.org/10.1016/j.indcrop.2011.04.018.
64. Reis, A. L. S., Damilano, E. D., Menezes, R. S. C., & De Morais, Jr. M. A., (2016).
Second-generation ethanol from sugarcane and sweet sorghum bagasse using the
yeast Dekkera bruxellensis. Ind. Crops Prod., 92, 255–262. https://doi.org/10.1016/j.
indcrop.2016.08.007.
65. Antonopoulou, I., Spanopoulos, A., & Matsakas, L., (2020). Single-cell oil and ethanol
production by the oleaginous yeast Trichosporon fermentans utilizing dried sweet
sorghum stalks. Renew. Energy, 146, 1609–1617. https://doi.org/10.1016/j.renene.
2019.07.107.
66. Wang, W., Zhuang, X., Yuan, Z., Yu, Q., Qi, W., Wang, Q., & Tan, X., (2012). High consis-
tency enzymatic saccharification of sweet sorghum bagasse pretreated with liquid hot
water. Bioresour. Technol., 108, 252–257. https://doi.org/10.1016/j.biortech.2011.12.092.
67. Kurian, J. K., Minu, K. A., Banerji, A., & Kishore, V. V., (2010). Bioconversion of
hemicellulose hydrolysate of sweet sorghum bagasse to ethanol by using Pichia stipitis
NCIM 3497 and Debaryomyces hansenii sp. BioResources, 5, 2404–2416.
68. Castro, E., Nieves, I. U., Rondón, V., Sagues, W. J., Fernández-Sandoval, M. T., Yomano,
L. P., York, S. W., et al., (2017). Potential for ethanol production from different sorghum
cultivars. Ind. Crops Prod., 109, 367–373. https://doi.org/10.1016/j.indcrop.2017.08.050.
69. Damay, J., Boboescu, I. Z., Duret, X., Lalonde, O., & Lavoie, J. M., (2018). A novel
hybrid first and second generation hemicellulosic bioethanol production process through
steam treatment of dried sorghum biomass. Bioresour. Technol., 263, 103–111. https://
doi.org/10.1016/j.biortech.2018.04.045.
70. Yu, J., Zhong, J., Zhang, X., & Tan, T., (2010). Ethanol Production from H2SO3 -Steam-
Pretreated Fresh Sweet Sorghum Stem by Simultaneous Saccharification and Fermenta-
tion, 401–409. https://doi.org/10.1007/s12010-008-8333-x.
71. Massoud, M. I., & El-razek, A. M. A., (2011). Suitability of Sorghum bicolor L. stalks
and grains for bioproduction of ethanol. Ann. Agric. Sci., 56(2), 83–87. https://doi.org/
10.1016/j.aoas.2011.07.004.
72. Sipos, B., Réczey, J., & Somorai, Z., (2009). Sweet Sorghum as Feedstock for Ethanol
Production: Enzymatic Hydrolysis of Steam-Pretreated Bagasse, 151–162. https://doi.
org/10.1007/s12010-008-8423-9.
73. Li, J., Li, S., Han, B., Yu, M., Li, G., & Jiang, Y., (2013). A novel cost-effective
technology to convert sucrose and homocelluloses in sweet sorghum stalks into ethanol.
Biotechnol. Biofuels, 6(174), 1–12. https://doi.org/http://www.biotechnologyforbiofuels.
com/content/6/1/174.
Apple Academic Press
74. Kaur, P., Uppal, S. K., Dhir, C., Sharma, P., & Kaur, R., (2015). Comparative study of
chemical pretreatments and acid saccharification of bagasse of sugar crops for ethanol
production. Sugar Tech., 17(4), 412–417. https://doi.org/10.1007/s12355-014-0346-x.
75. Yu, M., Li, J., Chang, S., Du, R., Li, S., Zhang, L., Fan, G., et al., (2014). Optimization
of ethanol production from NaOH-pretreated solid-state fermented sweet sorghum
bagasse. Energies, 7(7), 4054–4067. https://doi.org/10.3390/en7074054.
Author Copy
Author Copy
ROSA M. RODRÍGUEZ-JASSO, and HÉCTOR RUIZ LEZA
Biorefinery Group, Food Research Department, Faculty of Chemistry
Sciences, Autonomous University of Coahuila, Saltillo–25280, Coahuila,
Mexico, Phone: (+52)-1-844-416-12-38,
E-mail: rrodriguezjasso@uadec.edu.mx (R. M. Rodríguez-Jasso),
hector_ruiz_leza@uadec.edu.mx (H. A. Ruiz)
ABSTRACT
13.1 INTRODUCTION
Author Copy
people are exposed to environments with high levels of pollutants in the air
and it is resulting in the death of 4.2 million people worldwide per year due
to stroke, heart disease, lung cancer, and chronic respiratory diseases [2].
Air pollution is derived from a complex mixture of particle matters
(PM), vapors, and gases that can be emitted in different ways, both natural
and synthetic. PM is generally classified into particles of 10 (PM10) and 2.5
(PM2.5) micrometers and is formed by the conglomerate of suspended solids,
humidity, and atmospheric conditions, PM10 particles tend to accumulate in
the nasal concavities and the respiratory tract superior. At the same time, the
finer particles (< PM2.5) can also be absorbed in the lower respiratory tract,
causing diseases such as lung cancer, ischemic heart disease, respiratory tract
infections, allergies, and type 2 diabetes due to saturation of suspended solid
particles in the environment [1, 3]. On the other hand, gases released into the
atmosphere have produced anormal amounts of CO2, CO, NO(x), CH4, which
in addition to decreasing air quality and deteriorating human health, is the
cause of climate change because these gases cause the greenhouse effect,
which is the main reason of the increase in the temperature of the earth. The
increase in air contamination is directly related to the excessive use of fossil
fuels and industrial growth [1, 4].
Based on this problem, efforts have been made to try to mitigate or control
climate change. One of these efforts is The Paris Agreement of the United
Nations, whose main objective is to regulate polluting emissions that cause the
greenhouse effect and to keep the increase in world temperature below 2°C,
limiting this increase to 1.5°C and this in a time frame that allows ecosystems
to adapt to climate change and enable sustainable economic progress [5].
The biorefinery concept fits as an emerging solution to this problem
because it seeks the substitution of hydrocarbon-produced energy compounds
for biofuels produced from renewable sources such as biomass, some of
the proposed fuels are bioethanol, biogas, biohydrogen, biodiesel, bio-oil,
the raw material is oil, since different products, including energy compounds,
are produced from a petroleum complex mixture through various stages,
however, given the non-renewable nature of this material and the environ-
mental problems that result from the excessive use of these products, make
the concept of biorefinery an attractive alternative, since using different types
of biomass, which can come from various sources and have meager costs due
to its high availability, and result in a wide variety of compounds with novel
Author Copy
applications in addition to bioenergetics [7].
In general, biomass is defined as all material from a living organism,
and its use in biorefineries has generated a classification as generations,
depending on the biomass used. Currently, there are four generations of
biomass, the first generation (1G) comprises those of biorefineries that use
food crops as fuels, due to their high content in sugars, starch, and natural
oils, resulting in the first-generation biofuels, this generation represents
a significant concern because it uses edible resources creating direct
competition between food and fuel production [8]. The second-generation
biorefinery (2G) includes biorefineries whose raw material is LCB, that
is, those that are composed of hemicellulose, cellulose, and lignin. These
biorefineries present significant advantages since the raw material is found
in large quantities and is usually the residue of agricultural, forestry, and
food industries, so it does not compete with the crops generated for the food
sector, and due to the main components of the matrix it is possible to produce
a wide variety of byproducts and biofuels [9]. The third-generation (3G)
is characterized by the use of aquatic biomass as raw material, specifically
microalgae and macroalgae, this alternative is very attractive since it does
not generate competition with arable land, it also has a high carbohydrate
content and it does not have the strong union of terrestrial biomasses, so
there is a great opportunity to produce a wide variety of different products
and biofuels [10]. The fourth-generation (4G) consists of microorganisms
and genetically modified plants to have high amounts of carbon for the
production of fuels and various biochemicals [8].
Finally, the use of biomass has become a sustainable alternative to the
future for the supply of renewable products. It generates an alternative to
non-renewable energy sources, promoting the growth of a biologically based
economy, of a sustainable nature, also affected circular bioeconomy [8].
BIOREFINERIES
Author Copy
function as the main structuring agent of plants [12, 13]. This matrix of
components is an opportunity for the creation of different compounds from
a single raw material, through different chemical, physical, biological, and
enzymatic processes [12]. Finally, annual worldwide production of LCB of
200 × 109 tons per year is estimated, making abundant feedstocks available
for the production of biofuels and biochemicals [8, 14].
Some examples of possible products from the second generation
biorefineries are formic acid, ethylene glycol, acetic acid, lactic acid,
glycerol, glycolic acid from cellulose, but the primary designated use for this
polymer is for the production of bioethanol under the model of a biorefinery.
In the case of lignin, the transformation of this polymer can lead to quinones,
phenol benzene, syringaldehyde, pyruvate, and different lipids. And finally,
hemicellulose can be transformed into furfural, hydroxymethylfurfural (HMF),
levulinic acid, pentane, among others [15]. However, numerous investigations
have been directed towards the use of xylan from hemicellulose, because the
compounds derived from this polymer have various food applications such
as xylitol and xylose (Low-calorie sweetener substitutes) [13, 16], but with a
particular interest in the production of oligomers such as xylooligosaccharides
(XOs), which have multiple uses in food and pharmaceutical technology due
to their prebiotic, antioxidant, and cytoprotective activity [17], especially with
those of lower molecular weight or with a lower degree of polymerization
(DP = 2–4, oligomers made up of 2 to 4 xylose units) [13, 16] and those with
higher molecular weight, such as biopolymer with (DP > 5–10) which have
a wide variety of applications such as its use in the development of food
covers [18]. Finally, the design of operational chains that promote an integral
use of biomass, through the combination of the production of biofuels and
high added value co-products, because all these biochemicals promote the
economic projections of the second generation biorefineries. For this reason,
Author Copy
bioethanol using first-generation raw materials, due to its high content of
sugars, starches, or natural oils [22]. As reported by the Renewable Fuels
Association, 15,776 million bioethanol gallons were produced in 2019
using primarily grains like corn and sorghum [23]. However, although
first-generation biofuels diversify the raw materials with which biofuels are
produced, they are surrounded by environmental and social contradictions
such as environmental degradation due to the change in land use derived
from the expansion of agriculture, affecting indirectly or directly biodiversity
and the amount of CO2 produced, and also the creation of direct competition
with agriculture for food generation [22].
In addition, the environmental problem tends to worsen, because as the
population grows, there will be a higher demand for resources, specifically
energy and food, by 2030 it is expected that there will be 8.5 billion people
in the world, at a growth rate 83 million people each year, which projects an
increase of 70 million ha to meet global food demand by 2050 [24]. All these
factors drive towards a transition that first generates a renewable alterna-
tive for the production of fuels and does not compromise the use of food
resources and the environment, taking into account the raw material and the
sustainability of the process. Finally, this promotes the concept of second-
generation biorefineries, whose main objective is to generate bioeconomic
circular routes (Figure 13.1) where non-edible lignocellulosic waste is used
with increasingly optimal and sustainable processes [22, 24].
for agricultural production, and also is a sunny country, which also has various
seas, rivers, and lakes. For these reasons, Mexico has enormous potential to
develop more sustainable alternative energies such as solar, wind, hydraulic,
and thermal energy, but also with large amounts of biomass available due to
Apple Academic Press
Author Copy
FIGURE 13.1 Overview of second-generation biorefinery and its applications.
Mexico is still highly inclined towards the generation of energy through non-
renewable sources and where the burning of fossil fuels has a great presence
in the country [25].
Today, despite political limitations and the legal framework for the
Apple Academic Press
use of biomass for the production of biofuels is still not defined, there
has been a growing interest in the development of renewable energy in
the country. In 2016 the Mexican Center for Innovation in Bioenergy
(CEMIE-Bio) was created to carry out research and development of
technologies for the sustainable production of fuels from biomass, said
the group works in 5 specialized groups for each type of biofuel: solid
biofuels, bioalcohol, biodiesel, biogas, and bioturbosin. The results of
Author Copy
these research groups are expected to increase the use of biomass in
Mexico and contribute to the generation of at least 35% of clean energy
by 2024 [27].
The use of agave has a strong presence in Mexico since it is used in various
artisanal and industrial activities, this generates large amounts of solid waste
with enormous potential for its use [28]. Various of agave species contribute
to the economy of Mexico via the production of contrasting products such
as in northern Mexico, where agave is used for the production of textile
fibers, while, in the southern region tradition Mexico developing tequila,
mezcal, and pulque [28]. For example, Agave tequila weber is distinguished
for the production of tequila in central Mexico through cooking and crushing
this plant to create juices with high sugar content, which are fermented and
distilled to create tequila; however, this process leaves large amounts of solid
residue that can be used for the creation of biofuels and high added-value
compounds [29, 30].
Through the biorefinery model this solid bagasse can be used to
produce the biofuels and biproducts. Agave bagasse contains the cellu-
lose, hemicellulose, and lignin in (31–43%), (11–22%), and (11–20%)
w/w respectively. Biofuel refineries using the agave bagasse with the
maximum concentration of amorphous cellulose and hemicellulose, but
the recalcitrance property of lignin make it difficult to extract which act
as a barrier [31].
In addition to the presence of the Agave tequilana weber, the climate of
Mexico allows different species of Agave to grow such as Agave americana,
Agave angustifolia, Agave fourcroydes and Agave sisalana, and these also
form in the economic activity of the country [32]. Agave species reflects
the low rate of transpiration, have crassulacean acid metabolism capability
which upgrade water using efficiency in semiarid region [33].
Apple Academic Press
The Agave americana has broad leaves, and its color is green with slightly
gray with a whitish color, and also, is commonly used as an ornamental
plant in gardening or for the extraction of mead also known as “Aguamiel.”
Author Copy
These plants have very resistant fibers, used to make handcraft textile items.
Agave americana has an average size of 1.0 m to 1.4 m, with leaves 80 to
120 cm long and 15 to 20 cm wide, and its leaves are fleshy and smooth to
the touch [32].
The origin of this plant is Southern Africa and it is reported that the
size of this plant could reach a height of up to 2 meters. In Mexico, the
Agave americana L. is used in various activities, which makes a plant with
great importance for the economy and culture of the country. In Mexico,
generates the sustenance of 38,000 families, which also maintains a product
range in the country of 412,900 tons to 998,400 tons and wherein 2008 there
was a maximum in the production of this input with 1,125,100 tons, which
shows that it is a company that promises and has remained within Mexican
culture [34].
The Industrial Scientific Research Council (CSIR) studied about the
processing of Agave americana L. and highlighted the enormous potential
for fiber production and paper manufacturing, and also conducted important
market research on the global demand for inulin as a by-product which
could be derived from respected processes [34]. All this to carry out the
development of the technology to establish an industry based on the Agave
americana L. in Southern Africa, although studies have also been carried
out in our country to use in a more optimal way this plant and that can also
adapt to the process of manufacturing textile fibers since according to the
CSIR, the contribution of fiber in this plant is 1.5% resulting in the rest of
the plant having no use and also with these conditions prevents the exclusive
processing from this raw material to be unreachable. For the production of
mead Agave americana, L. is cut and generates a hole at the center of the
plant, and allow to collect the liquid of the plant also from the rain later it
could be supplied to tequila industries for the purification and sterilization
of the mead [34].
STAGE
The first stage contemplated for the biomass transformation process consists
of a series of mechanical treatments, which can be drying and crushing. This
process aims to improve the disposition of the material to the later stages of
the process. Generally, the dry and crushed materials have higher surface
areas and lower crystallinity, which can promote, for example, hydrolysis
Author Copy
or mass transfer phenomena within the biomass. Mechanical pretreatment
as the only pretreatment is not very useful for LCB es because it cannot
fractionate hemicellulose, cellulose, and lignin, so a series of subsequent
steps are necessary which can separate the biomass. However, mechanical
pretreatment is required to achieve essential characteristics in the process,
such as the appropriate particle size for later stages of the process, making
this an indispensable step for biorefineries [8].
The mechanical pretreatment of Agave Americana L consisted of a primary
stage where the leaves were cut into thin strips to facilitate drying. Subse-
quently, sunlight was used to dry the biomass partially, and then the strips were
cut in 1 cm2 pieces to be dried in a laboratory oven at 75°C until reaching an
internal humidity of 8% (w/w). Finally, a blade mill (Thomas Wiley, Swedes-
boro, NJ, USA) was used to achieve a particle size ≈ (1–0.3 mm) [36].
Only with the biomass, which had a particle size of (0.5–0.3 mm) the
characterization of the raw material was carried out. The analysis considered
were: a humidity determination at 121°C, ash determination, protein content
(N × 5.67) by Kjeldahl method, and solvent and aqueous extractives (through
a Soxhlet type extraction with acetone and water respectively). Also, it was
necessary to make a physicochemical characterization to determine the
cellulose, hemicellulose, and lignin content, which was made according to the
standard analytical procedures of the National Renewable Energy Laboratory
(NREL/TP-510-42618), through the quantification of monomers and acid by
high-performance liquid chromatography (HPLC) (Section 13.3.2.1). Finally,
Kalson lignin was quantified by the gravimetry method (Table 13.1) [36, 37].
TABLE 13.1 Characterization of Agave americana L. Leaves (% on Total Dry Weight) [38, 36]
Component Agave americana Agave tequilana Agave tequilana
L. Leaves [38] Bagasse [36]
Cellulose 29.89 ± 3.80 34.81 20.85
Apple Academic Press
Author Copy
Solvent extractives 2.55 ± 0.10 – 1.53
Aqueous extractives 43.55 ± 1.21 12.48 8.36
Because drying and grinding treatments are not capable of breaking the
complex union that exists between the components of lignocellulosic
material, it is necessary to add a stage to the process that has this task
as the main function. This stage is known as pretreatment. This opera-
tion is crucial since it directly influences the subsequent steps because it
also changes the adaptability of the raw material, such as for enzymatic
Apple Academic Press
Author Copy
and high-temperature pyrolysis. There are also chemical pretreatments
such as acid and alkaline treatments, oxidative treatments, ionic liquid,
Organosolv, among others. Another branch of pretreatments is made up
of biological processes carried out by microorganisms to structurally
change biomass and, finally, physicochemical treatments, in which there
are processes such as steam or CO2 explosion and finally hydrothermal
pretreatments [41].
Author Copy
metals. The Autohydrolysis technique reflects the economic feasibility
because the process does not require the acid, neutralizing agents, expensive
corrosive metals for the support of degradation of biomass, and prevention
from corrosion respectively [21, 42, 44].
The solid-liquid ratio of biomass and water composes the medium
complex and hike the processing scale which requires a greater amount of
the power consumption to heat the slurry mixture. Recent studies use the
heat exchanger to recover these amounts of energy which could enhance
the overall cost of the process because of the use of heat exchanger [45,
46]. Autohydrolysis technique adequate to improve the quality of solid
also removal of the metal component from the lignocellulosic feedstock.
Mainly removal of metal from the lignocellulosic raw material plays
a significant role during the production of bio-oil [47]. Studies state the
preference of the autohydrolysis technique reduces the chances of develop-
ment of inhibitors that would not disrupt the downstream processing and
complement its accessibility for efficient saccharification using hydrolyzes
enzymes [45, 48].
where; T is the process temperature (°C); t is the reaction time (min); 100
is the reference temperature in degree Celsius (°C); and w is an empirical
parameter related to the activation energy, assuming kinetics of first-order
(w=14.74) [50]. However, this equation in isothermal regime considers the
Apple Academic Press
Author Copy
When the hydrothermal pretreatment is in an isothermal regime, it is
necessary to make a distinction in the stages of the temperature profile. In
the case of the second equation, the first term refers to the heating stage,
so the value of t1 is the time in which the temperature setpoint is reached.
The isothermal period begins, which is described by the second term of the
equation and is evaluated from t1 to t2. With this temperature t2, the cooling
stage begins and ends with the final time of the process tf [36].
A central composite was used to identify the conditions in which the highest
Apple Academic Press
cellulose content was found in the pre-treated solid phase. To determine the
number of conditions was calculated using the following equation [36, 51].
N = 2k + 2 * k + 1 (4)
where; k is equal to the number of variables to manipulate during the experi-
ment (k = 2, time, and temperature). Resulting in a total of 9 conditions with
3 extra repetitions of the central value (170°C–30 min) [36, 51].
Author Copy
13.4.3.3 CHANGES IN BIOMASS DUE TO HYDROTHERMAL
PRETREATMENT
Finally, after the operation, 2 streams were generated, the first a solid phase
with a high concentration of cellulose and lignin with which the ethanol will
be produced, and a liquid one where the presence of XOs and monomers
will be checked during the treatment with the HPLC (Section 13.3.2.1) [51].
During the pretreatment stage, the heating time, isothermal, and cooling
time was captured each 10C (Figure 13.2). With this information, it was
possible to make a temperature profile and calculate the area under the curve.
The severity factor was calculated with the software Polymath v6.0., between
100°C of heating and 100°C of the cooling stage, for each pretreatment to
obtain the severity factor (Log(R0)) (Figure 13.3).
From biomass pretreatment and the calculation of the severity factor, it
was possible to measure the contribution of the operation in the transforma-
tion of the biomass and to relate this change to the increase in the severity
factor, as the pretreatment conditions were more intense. This factor is an
interesting relationship between the parameters of the pretreatment of LHW,
serving this same as a guideline to relate the behavior of biomass in different
reactors or the scaling-up of the processes (Table 13.2).
With the results of the previous table, we can see how hemicellulose
dilutes as the severity of the treatment increases. Furthermore, it is possible
to observe an increase in the percentage of cellulose and lignin in the solid
phase, starting from 36.67% to 50.66% (w/w) in the case of cellulose, and for
lignin from 18.38% to a 33.06% (w/w).
Besides, it is possible to see the reduction of factors such as pH, checking
the acidification of the medium by acetylation of the medium. XOs are an
resulting in some acids, such as acetic acid and degradation compounds such
as (HMF) or furfural [52].
Author Copy
It was determined that the best treatment for ethanol production, after
data analysis, is the condition with higher cellulose production and lower
hemicellulose content. For this reason, the state of 190°C and 50 min was
directly chosen for enzymatic hydrolysis, due to avoid inhibition of the
enzyme by traces of compounds derived from hemicellulose.
This behavior can be compared to that obtained by Aguirre-Fierro et
al. [53] where they used agave bagasse for ethanol production, which was
FIGURE 13.3 Glucose production from enzymatic hydrolysis condition. Author Copy
13.5 ENZYME SACCHARIFICATION OF AGAVE AMERICANA L. WASTE
Xylose 4.602±0.00 4.754±0.01 5.196±0.14 6.648±0.00 6.543±0.01 4.105±0.02 6.147±0.01 4.366±0.01 3.234±0.11
Arabinose 1.166±0.04 1.140±0.00 1.193±0.21 1.387±0.00 1.082±0.00 0.000±0.00 0.000±0.00 0.000±0.00 0.000±0.00
XOs 2.581 2.787 3.221 4.349 3.938 1.313 4.17 2.409 1.475
Acetic acid 0.117 0.129 0.144 0.124 0.126 0.128 0.157 0.185 0.215
Levulinic acid 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
Formic acid 0.000 0.000 0.000 0.000 0.000 0.025 0.253 0.034 0.035
Author Copy
388 Bioethanol: Biochemistry and Biotechnological Advances
Author Copy
such as α-arabinofuranosidase, ferulic acid esterase, acetyl xylan esterase,
and α-glucuronidase. Finally, the most common enzymes for cellulose
degradation are endo-glucanases, exo-glucanases, cellobiohydrolases, and
β-glucosidases; these enzymes can be found as a set for the hydrolysis of
cellulose to glucose [55].
The enzymatic cocktail used for the conversion of Agave americana L to biomass
was the cellulase Cellic® CTec2 by Novozymes (North America, USA). This
cocktail is composed of Exo-1,4-β-D-glucanase, Endo-1,4-β-D-glucanase, and
pH 4.8. Finally, reducing sugars were quantified with Miller’s method (DNS),
resulting in 123 FPU/mL [52, 56, 58].
Author Copy
calculated [56].
Author Copy
FIGURE 13.4 Graphical representation of the enzyme cocktail.
TABLE 13.3 Single-Factor Analysis of Variance for Yield and Initial Reaction Rate
Effect Degrees of Sum of Mean Sum F-Value Significant
Freedom Square of Square (p<0.005)
ANOVA-Enzymatic Saccharification Yield
Treatment 5 2257.729 451.546 206.153 F5,6 = 4.39
Error 6 13.142 2.19 – –
Total 11 2270.871 – – –
ANOVA-Initial Reaction Rates
Treatment 5 7.53 1.506 408.436 F5,6 = 4.39
Error 6 0.022 0.004 – –
Total 11 7.55 – – –
TABLE 13.4 Enzymatic Conversion Yield, Initial Rates and Mean Comparison of the Hydrolysis
Hydrolysis Enzymatic Comparison of Initial Rate of Comparison
Conditions Saccharification Means (95%) Reaction of Means
Yield (%) (g L–1 h–1) (95%)
Apple Academic Press
Author Copy
These results can be compared with the information reported by Láinez et al.
[59], who from Agave Salmiana leaves residues produced bioethanol using the
model of a second-generation biorefinery. Through an acid-alkaline pretreat-
ment, where the hemicellulose was removed, and enzymatic hydrolysis with a
solid load was 5.62% and 10 FPU/g of glucan, it resulted in 49.04 g/L of glucose
released and also a behavior similar to leading in the enzymatic hydrolysis of
Agave Americana L. The enzyme used during this experiment was Celluclast
1.5L (Novozymes), an enzymatic complex of cellulases and β-glucosidase.
According to the results obtained from the two previous stages, the isothermal
hydrothermal pretreatment of 190°C for 50 min, and the enzymatic hydrolysis
Author Copy
samples were taken every 12, 24, 36, and 48 hours to monitor fermentation.
Each of the tests was carried out in triplicate [52]. The conversion of glucose
to ethanol was measured as described by (Section 13.3.2.1) using HPLC.
The yeast strain PE-2 was grown during this experiment in a 500 mL flask,
with a total working volume of 125 mL. The medium contained 10 g/L
yeast extract, 10 g/L peptone, 10 g/L agar, 10 g/L dextrose, and 0.5 g/L of
(NH4)2HPO4 and MgSO4·7H2O. The glucose was dissolved in an aqueous
solution and sterilized separately. The flask with the medium and the yeast
was incubated at 30°C, 150 rpm for 12 h. The yeast suspension was made
from the pre-inoculum centrifugation at 5,000 rpm at 4°C for 10 min. Finally,
it was suspended in a 0.4% NaCl solution [35].
where; [EtOH] is the final concentration of ethanol (g/L); 0.51 is the theoretical
conversion of glucose to ethanol based on stoichiometric biochemistry
of yeast; [Biomass] is dry biomass concentration at the beginning of the
enzymatic saccharification (g/L); f is the cellulose fraction in dry biomass
(g/g); and 1.111 is the conversion factor of cellulose to glucose.
From the operation route outlined during this article, it was possible to
Apple Academic Press
Author Copy
ethanol by microorganisms as a carbon source.
of fossil fuels implies. From this study, it was observed that the remnants
of different processes, such as the use of Agave americana L, are an option
for the production of biofuels such as bioethanol, revaluing the residues of
this economic activity. Also, the use of waste from desert biomass implies
Apple Academic Press
generating lignocellulosic materials with less use of water and cares in their
production. For this and other reasons, it is necessary to continue in the
search for new technological alternatives that reduce the cost and increase the
production volumes of fuels such as bioethanol as well as it is also necessary
to generate options that integrate sustainable processes such as the use of
enzymes and hydroelectric pretreatment. Finally, the use of the residues of
Agave americana L, reinforces the concept of adaptability of biorefineries
Author Copy
and gives rise to a new area of application of resources that avoid competition
with food crops, an essential factor in developing countries such as Mexico.
ACKNOWLEDGMENTS
KEYWORDS
• biofuels
• biomass
• enzymatic hydrolysis
• fermentation
• hydrothermal processing
• lignocellulosic material
• pretreatment
• second-generation biorefinery
REFERENCES
1. Duan, R., Hao, K., & Yang, T., (2020). Air pollution and chronic obstructive pulmonary
disease. Chronic Dis. Transl. Med. https://doi.org/10.1016/j.cdtm.2020.05.004.
Apple Academic Press
Author Copy
5. Gao, Y., Gao, X., & Zhang, X., (2017). The 2°C global temperature target and the
evolution of the long-term goal of addressing climate change—from the United Nations
framework convention on climate change to the Paris agreement. Engineering, 3(2),
272–278. https://doi.org/10.1016/J.ENG.2017.01.022.
6. Carrillo-Nieves, D., Rostro, A. M. J., De La Cruz, Q. R., Ruiz, H. A., Iqbal, H. M. N.,
& Parra-Saldívar, R., (2019). Current status and future trends of bioethanol production
from agro-industrial wastes in Mexico. Renew. Sustain. Energy Rev., 102, 63–74. https://
doi.org/10.1016/j.rser.2018.11.031.
7. Ruiz, H. A., Thomsen, M. H., & Trajano, H. L., (2017). Hydrothermal Processing in
Biorefineries. Springer International Publishing, Cham, Switzerland. https://www.springer.
com/gp/book/9783319564562 (accessed on 28 October 2021).
8. Kumar, B., Bhardwaj, N., Agrawal, K., Chaturvedi, V., & Verma, P., (2020). Current
perspective on pretreatment technologies using lignocellulosic biomass: An emerging
biorefinery concept. Fuel Process. Technol., 199, 106244. https://doi.org/10.1016/j.
fuproc.2019.106244.
9. Ajao, O., Marinova, M., Savadogo, O., & Paris, J., (2018). Hemicellulose based integrated
forest biorefineries: Implementation strategies. Ind. Crops Prod., 126, 250–260. https://
doi.org/10.1016/j.indcrop.2018.10.025.
10. Lara, A., Rodríguez-Jasso, R. M., Loredo-Treviño, A., Aguilar, C. N., Meyer, A. S., &
Ruiz, H. A., (2020). Enzymes in the third generation biorefinery for macroalgae biomass.
In: Biomass, Biofuels, Biochemicals (pp. 363–396). Elsevier. https://doi.org/10.1016/
B978-0-12-819820-9.00017-X.
11. Singh, A., Rodríguez, J. R. M., Gonzalez-Gloria, K. D., Rosales, M., Belmares, C. R.,
Aguilar, C. N., Singhania, R. R., & Ruiz, H. A., (2019). The enzyme biorefinery platform
for advanced biofuels production. Bioresour. Technol. Reports, 7, 100257. https://doi.
org/10.1016/j.biteb.2019.100257.
12. Kupski, L., Telles, A. C., Gonçalves, L. M., Nora, N. S., & Furlong, E. B., (2018). Recovery
of functional compounds from lignocellulosic material: An innovative enzymatic
approach. Food Biosci., 22, 26–31. https://doi.org/10.1016/j.fbio.2018.01.001.
13. Banerjee, S., Patti, A. F., Ranganathan, V., & Arora, A., (2019). Hemicellulose based biore-
finery from pineapple peel waste: Xylan extraction and its conversion into xylooligosac-
charides. Food Bioprod. Process, 117, 38–50. https://doi.org/10.1016/j.fbp.2019.06.012.
14. Haykiri-Acma, H., Yaman, S., & Kucukbayrak, S., (2010). Comparison of the thermal
reactivities of isolated lignin and holocellulose during pyrolysis. Fuel Process. Technol.,
91(7), 759–764. https://doi.org/https://doi.org/10.1016/B978-0-12-818178-2.00005-5.
15. Islam, M. K., Wang, H., Rehman, S., Dong, C., Hsu, H. Y., Lin, C. S. K., & Leu, S. Y.,
(2020). Sustainability metrics of pretreatment processes in a waste derived lignocellulosic
biomass biorefinery. Bioresour. Technol., 298, 122558. https://doi.org/10.1016/j.biortech.
2019.122558.
Apple Academic Press
16. Singh, R. D., Talekar, S., Muir, J., & Arora, A., (2019). Low degree of polymerization
xylooligosaccharides production from almond shell using immobilized nano-bio catalyst.
Enzyme Microb. Technol., 130, 109368. https://doi.org/10.1016/j.enzmictec.2019.109368.
17. Samanta, A. K., Jayapal, N., Jayaram, C., Roy, S., Kolte, A. P., Senani, S., & Sridhar, M.,
(2015). Xylooligosaccharides as prebiotics from agricultural by-products: Production and
applications. Bioact. Carbohydrates Diet. Fibre, 5(1), 62–71. https://doi.org/10.1016/j.
bcdf.2014.12.003.
18. Ruiz, H. A., Rodríguez-Jasso, R. M., Fernandes, B. D., Vicente, A. A., & Teixeira, J. A.,
(2013). Hydrothermal processing, as an alternative for upgrading agriculture residues
Author Copy
and marine biomass according to the biorefinery concept: A review. Renew. Sustain.
Energy Rev., 21, 35–51. https://doi.org/10.1016/j.rser.2012.11.069.
19. Galanopoulos, C., Giuliano, A., Barletta, D., & Zondervan, E., (2020). An integrated
methodology for the economic and environmental assessment of a biorefinery supply
chain. Chem. Eng. Res. Des., 160, 199–215. https://doi.org/10.1016/j.cherd.2020.05.016.
20. Ubando, A. T., Felix, C. B., & Chen, W. H., (2020). Biorefineries in circular bioeconomy:
A comprehensive review. Bioresour. Technol., 299. https://doi.org/10.1016/j.biortech.
2019.122585.
21. Ruiz, H. A., Vicente, A. A., & Teixeira, J. A., (2012). Kinetic modeling of enzymatic
saccharification using wheat straw pretreated under autohydrolysis and organosolv
process. Ind. Crops Prod., 36, 100–107. https://www.sciencedirect.com/science/article/
pii/S0926669011003682 (accessed on 28 October 2021).
22. Correa, D. F., Beyer, H. L., Possingham, H. P., Thomas-Hall, S. R., & Schenk, P. M., (2017).
Biodiversity impacts of bioenergy production: Microalgae vs. first-generation biofuels.
Renew. Sustain. Energy Rev., 74, 1131–1146. https://doi.org/10.1016/j.rser.2017.02.068.
23. Renewable Fuels Association. https://ethanolrfa.org/markets-and-statistics/view-all-
markets-and-statistics (accessed on 28 October 2021).
24. Hassan, S. S., Williams, G. A., & Jaiswal, A. K., (2019). Moving towards the Second
generation of lignocellulosic biorefineries in the EU: Drivers, challenges, and opportunities.
Renew. Sustain. Energy Rev., 101, 590–599. https://doi.org/10.1016/j.rser.2018.11.041.
25. Ruiz, H. A., Martínez, A., & Vermerris, W., (2016). Bioenergy potential, energy crops, and
biofuel production in Mexico. BioEnergy Res., 9(4), 981–984. https://doi.org/10.1007/
s12155–016–9802–7.
26. Secretary of Energy (SENER). (2020). National Energy Balance: Total Final Energy
Consumption by Fuel. https://sie.energia.gob.mx/bdiController.do?action=cuadro&
cvecua=IE7C01 (accessed on 28 October 2021).
27. Honorato-Salazar, J. A., & Sadhukhan, J., (2020). Annual biomass variation of agriculture
crops and forestry residues, and seasonality of crop residues for energy production in
Mexico. Food Bioprod. Process, 119, 1–19. https://doi.org/10.1016/j.fbp.2019.10.005.
28. Ponce-Ortega, J. M., & Santibañez-Aguilar, J. E., (2019). Optimization of the supply
chain associated to the production of bioethanol from residues of agave from the tequila
process in Mexico. In: Strategic Planning for the Sustainable Production of Biofuels
(pp. 115–146). Elsevier. https://doi.org/10.1016/B978-0-12-818178-2.00005-5.
29. Mielenz, J. R., (2001). Ethanol production from biomass: Technology and commer-
cialization status. Curr. Opin. Microbiol., 4(3), 324–329. https://doi.org/10.1016/S1369-
5274 (00)00211-3.
30. Perez-Pimienta, J. A., Flores-Gómez, C. A., Ruiz, H. A., Sathitsuksanoh, N., Balan, V.,
Apple Academic Press
da Costa, S. L., Dale, B. E., Singh, S., & Simmons, B. A., (2016). Evaluation of agave
bagasse recalcitrance using AFEXTM, autohydrolysis, and ionic liquid pretreatments.
Bioresour. Technol., 211, 216–223.
31. Palomo-Briones, R., López-Gutiérrez, I., Islas-Lugo, F., Galindo-Hernández, K. L.,
Munguía-Aguilar, D., Rincón-Pérez, J. A., Cortés-Carmona, M. Á., et al., (2018). Agave
bagasse biorefinery: Processing and perspectives. Clean Technol. Environ. Policy, 20(7),
1423–1441. https://doi.org/10.1007/s10098-017-1421-2.
32. Parra, N. L. A., Del, V. Q. P., & Prieto, R. A., (2010). Extraction of agave fibers to make
paper and crafts. Acta Univ., 20, 77–83. https://doi.org/10.15174/au.2010.63.
Author Copy
33. Valdez-Vazquez, I., Alatriste-Mondragón, F., Arreola-Vargas, J., Buitrón, G., Carrillo-
Reyes, J., León-Becerril, E., Mendez-Acosta, H. O., et al., (2020). A comparison
of biological, enzymatic, chemical and hydrothermal pretreatments for producing
biomethane from agave bagasse. Ind. Crops Prod., 145, 112160. https://doi.org/10.1016/j.
indcrop.2020.112160.
34. Ganduri, L., Van, D. M. A. F., & Matope, S., (2015). Economic model for the production
of spirit, inulin and syrup from the locally eco-friendly agave americana. Procedia
CIRP, 28, 173–178. https://doi.org/10.1016/j.procir.2015.04.030.
35. Pereira, F. B., Romaní, A., Ruiz, H. A., Teixeira, J. A., & Domingues, L., (2014). Industrial
robust yeast isolates with great potential for fermentation of lignocellulosic biomass.
Bioresour. Technol., 161, 192–199. https://doi.org/10.1016/j.biortech.2014.03.043.
36. Pino, M. S., Rodríguez-Jasso, R. M., Michelin, M., & Ruiz, H. A., (2019). Enhancement
and modeling of enzymatic hydrolysis on cellulose from agave bagasse hydrothermally
pretreated in a horizontal bioreactor. Carbohydr. Polym., 211, 349–359. https://doi.
org/10.1016/j.carbpol.2019.01.111.
37. Devi, R., & Sit, N., (2019). Effect of single and dual steps annealing in combination
with hydroxypropylation on physicochemical, functional and rheological properties
of barley starch. Int. J. Biol. Macromol., 129, 1006–1014. https://doi.org/10.1016/j.
ijbiomac.2019.02.104.
38. Rijal, D., Vancov, T., McIntosh, S., Ashwath, N., & Stanley, G. A., (2016). Process
options for conversion of Agave tequilana leaves into bioethanol. Ind. Crops Prod., 84,
263–272. https://doi.org/10.1016/j.indcrop.2016.02.011.
39. Corbin, K. R., Byrt, C. S., Bauer, S., DeBolt, S., Chambers, D., Holtum, J. A. M., Karem,
G., et al., (2015). Prospecting for energy-rich renewable raw materials: Agave leaf case
study. PLoS One, 10(8), e0135382. https://doi.org/10.1371/journal.pone.0135382.
40. Sindhu, R., Binod, P., & Pandey, A., (2016). Biological pretreatment of lignocellulosic
biomass:An overview. Bioresour. Technol., 199, 76–82. https://doi.org/10.1016/j.biortech.
2015.08.030.
41. Chen, H., Liu, J., Chang, X., Chen, D., Xue, Y., Liu, P., Lin, H., & Han, S., (2017). A
review on the pretreatment of lignocellulose for high-value chemicals. Fuel Process.
Technol., 160, 196–206. https://doi.org/10.1016/j.fuproc.2016.12.007.
42. Lara-Flores, A. A., Araújo, R. G., Rodríguez-Jasso, R. M., Aguedo, M., Aguilar, C. N.,
Trajano, H. L., & Ruiz, H. A., (2018). Bioeconomy and Biorefinery: Valorization of
Rodríguez, R., Teixeira, J. A., & Ruiz, H. A., (2018). Bioreactor design for enzymatic
hydrolysis of biomass under the biorefinery concept. Chem. Eng. J., 347, 119–136.
https://www.sciencedirect.com/science/article/pii/S1385894718306235 (accessed on
28 October 2021).
44. Ruiz, H. A., Conrad, M., Sun, S. N., Sanchez, A., Rocha, G. J. M., Romaní, A., Castro,
E., et al., (2020). Engineering aspects of hydrothermal pretreatment: From batch to
continuous operation, scale-up and pilot reactor under biorefinery concept. Bioresour.
Technol., 299, 122685. https://doi.org/10.1016/j.biortech.2019.122685.
45. Fujimoto, S., Inoue, S., & Yoshida, M., (2018). High solid concentrations during the
Author Copy
hydrothermal pretreatment of eucalyptus accelerate hemicellulose decomposition and
subsequent enzymatic glucose production. Bioresour. Technol. Reports, 4, 16–20.
https://doi.org/10.1016/j.biteb.2018.09.006.
46. Liu, W., Wu, R., Hu, Y., Ren, Q., Hou, Q., & Ni, Y., (2020). Improving enzymatic
hydrolysis of mechanically refined poplar branches with assistance of hydrothermal
and Fenton pretreatment. Bioresour. Technol., 316, 123920. https://doi.org/10.1016/j.
biortech.2020.123920.
47. Ge, J., Wu, Y., Han, Y., Qin, C., Nie, S., Liu, S., Wang, S., & Yao, S., (2020). Effect of
hydrothermal pretreatment on the demineralization and thermal degradation behavior
of eucalyptus. Bioresour. Technol., 307, 123246. https://doi.org/10.1016/j.biortech.
2020.123246.
48. Mariano, A. P. B., Unpaprom, Y., & Ramaraj, R., (2020). Hydrothermal pretreatment
and acid hydrolysis of coconut pulp residue for fermentable sugar production. Food
Bioprod. Process, 122, 31–40. https://doi.org/10.1016/j.fbp.2020.04.003.
49. Overend, R. P., Chornet, E., & Gascoigne, J. A., (1987). Fractionation of lignocellulosics
by steam-aqueous pretreatments. Philos. Trans. R. Soc. London. Ser. A, Math. Phys. Sci.,
321(1561), 523–536. https://doi.org/10.1098/rsta.1987.0029.
50. Ruiz, H. A., Cerqueira, M. A., Silva, H. D., Rodríguez-Jasso, R. M., Vicente, A. A.,
& Teixeira, J. A., (2013). Biorefinery valorization of autohydrolysis wheat straw
hemicellulose to be applied in a polymer-blend film. Carbohydr. Polym., 92(2),
2154–2162. https://doi.org/10.1016/j.carbpol.2012.11.054.
51. Ruiz, H. A., Silva, D. P., Ruzene, D. S., Lima, L. F., Vicente, A. A., & Teixeira, J.
A., (2012). Bioethanol production from hydrothermal pretreated wheat straw by a
flocculating Saccharomyces Cerevisiae strain - effect of process conditions. Fuel, 95,
528–536. https://doi.org/10.1016/j.fuel.2011.10.060.
52. Aguilar, D. L., Rodríguez-Jasso, R. M., Zanuso, E., De Rodríguez, D. J., Amaya-Delgado, L.,
Sanchez, A., & Ruiz, H. A., (2018). Scale-up and evaluation of hydrothermal pretreatment
in isothermal and non-isothermal regimen for bioethanol production using agave bagasse.
Bioresour. Technol., 263, 112–119. https://doi.org/10.1016/j.biortech.2018.04.100.
53. Aguirre-Fierro, A., Ruiz, H. A., Cerqueira, M. A., Ramos-González, R., Rodríguez-Jasso,
R. M., Marques, S., & Lukasik, R. M., (2020). Sustainable approach of high-pressure agave
bagasse pretreatment for ethanol production. Renew. Energy., 155, 1347–1354. https://
www.sciencedirect.com/science/article/pii/S0960148120305875 (accessed on 28 October
2021).
54. Pellis, A., Cantone, S., Ebert, C., & Gardossi, L., (2018). Evolving biocatalysis to meet
bioeconomy challenges and opportunities. N. Biotechnol., 40, 154–169. https://doi.org/
10.1016/j.nbt.2017.07.005.
55. Champreda, V., Mhuantong, W., Lekakarn, H., Bunterngsook, B., Kanokratana, P., Zhao,
Apple Academic Press
X. Q., Zhang, F., et al., (2019). Designing cellulolytic enzyme systems for biorefinery:
From nature to application. J. Biosci. Bioeng., 128(6), 637–654. https://doi.org/10.1016/j.
jbiosc.2019.05.007.
56. Adney, B., & Nrel, J. B., (2008). Measurement of Cellulase Activities. NERL/TP-510–
42628. National Renewable Energy Laboratory, Golden, CO.
57. Dowe, N., & Mcmillan, J., (2008). SSF Experimental Protocols--Lignocellulosic
Biomass Hydrolysis and Fermentation: Laboratory Analytical Procedure (LAP). Issue
Date: 10/30/2001.
58. Rodrigues, A. C., Haven, M. Ø., Lindedam, J., Felby, C., & Gama, M., (2015). Celluclast
Author Copy
and cellic® CTec2: Saccharification/fermentation of wheat straw, solid-liquid partition
and potential of enzyme recycling by alkaline washing. Enzyme Microb. Technol., 79, 80,
70–77. https://doi.org/10.1016/j.enzmictec.2015.06.019.
59. Láinez, M., Ruiz, H. A., Arellano-Plaza, M., & Martínez-Hernández, S., (2019). Bioeth-
anol production from enzymatic hydrolysates of Agave salmiana leaves comparing S.
cerevisiae and K. Marxianus. Renew. Energy, 138, 1127–1133. https://doi.org/10.1016/j.
renene. 2019.02.058.
Author Copy
OZNUR YILDIRIM, DOGUKAN TUNAY, BESTAMI OZKAYA, and
AHMET DEMIR
Department of Environmental Engineering, Yildiz Technical University,
Turkey, Davutpasa Campus, Istanbul–34220, Turkey
ABSTRACT
Author Copy
Graphical abstract
14.1 INTRODUCTION
In recent years, ethanol production from LCB has gained great momentum.
LCB mostly consists of agricultural wastes and forestry residues [1]. Energy
production from LCB is more significant in the economic, social, and environ-
mental aspects for all over the world. Agricultural wastes are substantial focus
area for renewable energy production because of containing high amounts
of carbohydrates. Special pretreatment applications should be applied to
make LCB open to biological activities by breaking the rigid structure of the
lignocellulosic matrix [2, 3].
Cellulose, a kind of polymer, found abundantly in nature, formed as a
result of glucose monomers combined by glucosidic bonds [4]. Cellulose
polymers of different lengths and structures bond to each other with weak
hydrogen bonds to form amorphous or crystalline cellulose fibrils [5].
As the number of cellulose chains formed increases, the strength of the
cellulose increases, and an indigestible crystal structure is formed, which
makes it arduous to break down. Similarly, as the Glucose number of cellu-
lose increases, it is assumed that the DP increases which is proportional
to the glucose number [6]. The most significant limiting factor affecting
the efficiency of hydrolysis is the cellulose crystallinity. The amorphous
cellulose chains can decompose faster than those with a crystalline struc-
ture [7]. Hemicellulose, which is formed by the combination of different
Author Copy
With the use of lignocellulosic ethanol in the transportation sector, the net
CO2 emissions released into the atmosphere are decreasing [14]. However,
it is no longer sufficient to produce the maximum amount of ethanol by
increasing lignocellulosic ethanol production efficiency with different
pretreatment processes. Due to the rapid reduction of water resources, water
usage during the process and its reduction have become one of the most
important parameters. It is known that most water usage in the lignocellulose-
based ethanol production occurs during the pretreatment stage [15]. Water
usage should be reduced in the pretreatment stage to make this process more
sustainable. It is anticipated that if the water resources in the world continue
to be consumed at the current rate, it will be able to manage for about 30
more years. Water demand is expected to increase by 40% by 2050, and 25%
of people will not have adequate access to freshwater [16].
For the production of ethanol, water is essential, especially in the
grinding, liquefaction, and fermentation processes. In recent years, water
consumption per gallon of ethanol has decreased significantly. According to
the plants that were operated in 2002, older plants consuming much more
water. For instance, water usage of older plants were more than the 15-Gal
water/Gal etOH. However, it was declined from 1 gallon to 11 gallons and
on average 4.7 gallons for production of 1 gallon of ethanol from the 2002
data (Figure 14.1) [17].
There are two types of fermentation processes that is used in the ethanol
industry, which are continuous and batch. 27% of ethanol was produced with
the continuous fermentation process, and the remaining was produced with
the batch process according to the 2002 survey of ethanol plants. When the
ethanol plants were analyzed due to their process selection, it can be said
that the continuous fermentation system was more preferred in large plants.
While 4 of these large plants were operated continuous, 17 out of 21 plants
were operated with batch system [17].
Author Copy
FIGURE 14.1 Gallons of water usage per gallon of ethanol produced for the 21 ethanol
plants in 2002 [17].
green, and gray water footprints. Green water footprint can be explained as
the water that is deposited, transformed, evaporated by plants, or consump-
tion of water from precipitation that is held by the roots of the soil. In short,
everything is incorporated with plant, agriculture, and forestry taking charge
Apple Academic Press
of water cycle called the green water footprint. On the other hand, surface
and groundwater sources that are used for domestic, industrial, and irriga-
tion water for agricultural purposes are related to the blue water footprint.
Bluewater consumption estimations need detailed analyzes and study since
there could be many reasons for the water usage directly or indirectly. For
instance, there is a study regarding the water footprint for coffee and tea
consumption in the Netherlands. If the virtual and real water consumptions
Author Copy
were compared, there will be a huge difference. For example, while the real
water consumptions were 0.125 L and 0.250 L water for the coffee and tea,
respectively. On the other hand, virtual water consumptions rise to approxi-
mately 140 L and 34 L for coffee and tea [19]. Greywater footprint is the
other component that is considered the freshwater usage for the assimilation
of pollutants according to the water quality standards. It is considered as a
point source pollution discharge through a pipeline, runoff, or impervious
surfaces.
Author Copy
production stage and irrigation need for different locations of the United
States. Bioethanol production consumes 0.5–28% of the total water which
includes process water and irrigation water. It shows that growing crops for
bioethanol production has a much higher water footprint due to the lignocel-
lulosic waste or uncontrolled growing herbs or plants usage.
FIGURE 14.2 The components of water footprint during biofuel production from lignocel-
lulosic biomass.
the amount of water consumed during the pretreatment, in other words, the
determination of the blue water footprint for pretreatment will be examined
in current studies.
Author Copy
The carbohydrate polymers in lignocellulose are tightly bonded to lignin
with hydrogen bonds and covalent bonds, creating a complex and crystalline
structure that prevents the carbohydrate in the raw material from being used
by microorganisms and enzymes [28]. The object of the pretreatment is to
decrease the crystallinity of lignocellulose, break their resistance to enzymatic
depolymerization, and disrupt the heterogeneous structure of lignocellulosic
materials. For obtaining valuable products from lignocellulosic materials,
their complex structure must be disrupted and broken down into smaller
carbohydrate components. Pretreatment types are basically classified as
physical, chemical, physicochemical, and biological.
For a conventional pretreatment process to be efficient; It would be
expected to increase the accessible cellulose surface area, break the lignin
barrier and cellulose crystallinity, prevent the formation of toxic by-products
that will cause inhibition in the fermentation stages, to reduce the loss of
reducing sugar resulting from the process and to be economically efficient
[29]. However, new issues to be considered in addition to these items have
become a current issue. Now, in addition to the requirements such as high
sugar conversion rate of the pretreatment process, being economical, and
providing low inhibitor production, lower water usage has become one of the
most important parameters. Due to the depletion of water resources and GW,
efforts to develop new methods that reduce water consumption during the
process and wastewater production after the process to make environmen-
tally friendly waste-to-fuel production processes such as biofuel production
from lignocellulosic wastes more sustainable.
The literature includes different approaches for pretreatment. Several
methods of pretreatment were tested by applying them to different lignocel-
lulosic wastes. In this chapter, studies that mostly conduct water footprint
studies and indicate how much water is consumed per waste are included.
reducing the size of wastes by cutting, grinding, and milling [30]. Methods
such as mechanical extrusion, grinding, and microwave are in the physical
pretreatment category. In addition, there are methods such as ultrasound and
pulsed electric field, which have recently been applied.
In milling and chipping methods, which are the most preferred types of
physical treatment, size reduction is performed with the help of grinders and
cutting tools without any water usage. However, it may be necessary to use
Author Copy
water in some methods such as microwave. One of the physical processes
that are used for biofuel production from lignocellulosic feedstocks is the
cooling necessity. Wu and Sawyer [31] found that 65% of the total process
of water consumption for renewable diesel blendstock (RDB) is caused by
cooling. Water necessity for the pretreatment and enzymatic hydrolysis were
found 0.16 L water/L RDB and 0.18 L water/L RDB, respectively.
acid and base, the washing is a significant step to bring the pH to a neutral
level. Approximately 10 L of water is consumed during the 1 L of ethanol
production process [35].
In a study [36], a greenhouse gas (GHG) and water use inventory were
Apple Academic Press
Author Copy
compared in terms of water consumption, WW and iHG processes have been
observed to have almost similar water consumption (about 0.30–0.40 L/MJ).
The reason the water usage values are similar for the WW process and the
iHG process is that all the water used in the WW stage is eventually recovered
and recycled through highly efficient evaporative dehydration process.
According to the National Renewable Energy Laboratory Technical
Report, Aden et al. [37] investigated the ethanol process design and econom-
ical approaches for corn stover biomass. Water balances were derived from
the ASPEN simulations. It was found that 0.252 kg ethanol can be produced
from 1 kg of corn stover with the 0.224 kg of water consumption for the
pre-hydrolysis (0.113 kg of water/kg ethanol produced) and saccharification
(0.111 kg of water/kg ethanol produced). According to ASPEN simulations,
it was proven the pretreatment applied during ethanol generation from the
corn stove affects the water usage by 2.62%.
The processes where physical and chemical pretreatment processes are used
together are defined as a physicochemical pretreatment method [2]. In the
methods applied in this type of treatment, treatment is done in harsh condi-
tions such as high temperature and pressure. LHW, wet oxidation, ammonia
fiber explosion (AFEX), and CO2 explosion are the most used techniques.
As the name suggests, methods such as LHW and wet oxidation are methods
that cause too much water consumption. In terms of process efficiency, they
have less sugar conversion efficiency compared to acid and base applications
under high temperature and pressure.
Dong et al. [38], produced ethanol using sulfite pretreated Monterey pine
and completed the water and energy footprint analysis of the whole process.
As a new pretreatment in the study, the process of breaking down oven-
dried (OD) wood chips with sulfuric acid and sodium bisulfite was applied
Apple Academic Press
in a rotary electric heated digester. They proved that the newly pretreatment
process they performed when they examined the water footprint of the
process provides 3.65 tons/ton of dry biomass water consumption and this
value provides 25–51% less water consumption than pretreatment processes
such as steam explosion (SE). They also claimed in their report that the
highest water consumption of the SE, diluted acid, and organosolv pretreat-
ment types where organosolv > dilute acid > steam explosion, respectively.
Author Copy
Organosolv pretreatment leads not only to the highest water consumption,
but also to energy consumption. Pretreatment with dilute acid consumes less
energy than steam putting process. It is also stated that the acid pretreatment
process mostly results in higher sugar conversion efficiency than SE [39]. In
this case, if both the efficiencies of energy, water consumption, and pretreat-
ment are compared, the most preferred pretreatment form, dilute acid, tends
to be more advantageous.
In another study [40], while investigating the effect of reactor filling and
solid-liquid ratio on ethanol yield during the production of ethanol from corn
stover, water, and steam consumption amounts were investigated. At the end
of the process, it was observed that the steam consumption was significantly
reduced and no acidic waste containing acid was produced. in pretreatment
conditions where the highest sugar content and minimum water consumption
occur; 190°C; 3.00% sulfuric acid; 3 min, water, and steam consumption,
respectively, 97.3 g water and 44 g steam per g dry waste were used.
Pan et al. [41], after dilute acid pretreatment for corn stover based ethanol
production, water footprint analysis of various conditioning processes were
carried out. These methods were; overliming, ammonia addition, two-stage
treatment, and membrane separations. In overliming with conditioning
application, 4.94 kg/kg dry biomass direct process water was used in total
and approximately 35% of this was used during pretreatment. In conditioning
with ammonia and two-stage conditioning application, a total of 4.58 kg and
4.25 kg of water were used respectively, and the most water was used in the
pH adjustment stage. In the air conditioning application with membrane, the
total water usage (2.54 kg) has the lowest value compared to other processes.
The results showed that the amount of steam and water used in the dilute acid
pretreatment stage ranged from 14% to 74% for the four methods applied.
The 4.94 L of water per kg-DB was used in conventional overliming. With
Author Copy
14.7 WATER CONSUMPTION IN BIOLOGICAL PRE-TREATMENT
solid residue to 80–90%. In addition, the amount of water used is saved about
10 to 20 times the weight of the cellulose solid residue, which significantly
reduces the water used in production, resulting in an important decrease in
the water footprint. As in the mentioned study, the original efficiency of the
process is preserved and even increased, while water consumption can be
reduced by new methods (Table 14.1).
As a result, there are many pretreatment methods used for lignocellulose-
Author Copy
based ethanol production. The sugar conversion efficiency of the pretreatment
method used varies according to the type, content, location, and harvesting
method. In order to figure out which pretreatment process would result
in maximum efficiency, different parameters such as the amount of water
consumption and output in the production of ethanol need to be examined.
Table 14.1 shows the pretreatment type applied, the LCB used, the amount of
water needed for ethanol production, and the ethanol production efficiency
of some studies.
14.8 CONCLUSION
Pretreatment
Ammonia fiber explosion Corn stover 1 N/A [51, 55]
Soaking in aqueous ammonia N/A 5.53 0.148 [45, 46]
(SAA)
Biochemical conversion (diluted Corn stover 8.55 0.252 [56]
Author Copy
414 Bioethanol: Biochemistry and Biotechnological Advances
all aspects. By developing new methods that will reduce the high freshwater
consumption, which is one of the handicaps of this process, the process will
be made more preferred. More novel technology development and techno-
economic analysis of these methods are required for lignocellulosic ethanol
Apple Academic Press
production.
KEYWORDS
• biorefinery
Author Copy
• ethanol
• ionic liquids
• lignocellulosic biomass
• pretreatment
• water footprint
REFERENCES
1. Anwar, Z., Gulfraz, M., & Irshad, M., (2014). Agro-industrial lignocellulosic biomass a key
to unlock the future bio-energy: A brief review. J. Radiat. Res. Appl. Sci., 7(2), 163–173.
2. Batista, M. D., De Oca-Vásquez, G., Vega-Baudrit, J. R., Rojas-Álvarez, M., Corrales-
Castillo, J., & Murillo-Araya, L. C., (2020). Pretreatment methods of lignocellulosic
wastes into value-added products: Recent advances and possibilities. Biomass Convers.
Biorefinery, 1–18.
3. Mancini, G., Papirio, S., Lens, P. N. L., & Esposito, G., (2016). Solvent pretreatments
of lignocellulosic materials to enhance biogas production: A review. Energy & Fuels,
30(3), 1892–1903.
4. Wang, B. T., Hu, S., Yu, X. Y., Jin, L., Zhu, Y. J., & Jin, F. J., (2020). Studies of cellulose
and starch utilization and the regulatory mechanisms of related enzymes in Fungi.
Polymers (Basel), 12(3), 1–17.
5. Chundawat, S. P. S., et al., (2011). Restructuring the crystalline cellulose hydrogen bond
network enhances its depolymerization rate. J. Am. Chem. Soc., 133(29), 11163–11174.
6. Zoghlami, A., & Paës, G., (2019). Lignocellulosic biomass: Understanding recalcitrance
and predicting hydrolysis. Front. Chem., 7.
7. Yu, Y., et al., (2010). Significant differences in the hydrolysis behavior of amorphous
and crystalline portions within microcrystalline cellulose in hot-compressed water. Ind.
Eng. Chem. Res., 49(8), 3902–3909.
8. Angell, S., Norris, F. W., & Resch, C. E., (1936). The analysis of carbohydrates of the
cell wall of plants: The determination of pentoses as single substances and in mixtures
containing uronic acids and hexoses. Biochem. J., 30(12), 2146–2154.
Author Copy
201–224). Elsevier.
14. ICF, (2017). A life-cycle analysis of the greenhouse gas emissions of corn-based ethanol.
Ind. Biotechnol., 13(1), 19–22.
15. Karimi, K., (2015). Lignocellulose-Based Bioproducts. Switzerland: Springer Interna-
tional Publishing. (p. 270).
16. Shaikh, A., (2017). The Bad News? The World will Begin Running Out of Water By 2050.
The good news? It’s not 2050 yet. United Nations News & Commentary Global News.
17. Shapouri, H., Gallagher, P., & Graboski, M. S. (2002). USDA's 1998 ethanol cost-of-
production survey (No. 1473–2021–024).
18. Gleick, P. H., et al., (1993). World freshwater resources. Water Cris. a Guide. to World's
Freshwater Resour.
19. Chapagain, A. K., & Hoekstra, A. Y., (2007). The water footprint of coffee and tea
consumption in the Netherlands. Ecol. Econ., 64(1), 109–118.
20. De Jong, E., & Jungmeier, G., (2015). Biorefinery Concepts in Comparison to Petro-
chemical Refineries. Elsevier B.V.
21. Hingsamer, M., & Jungmeier, G., (2018). Biorefineries. Elsevier Inc.
22. Gnansounou, E., & Pandey, A., (2017). Classification of Biorefineries Taking into
Account Sustainability Potentials and Flexibility. Elsevier B.V.
23. Council, N. R., et al., (2008). Water Implications of Biofuels Production in the United
States. National Academies Press.
24. Lingaraju, B. P., Lee, J. Y., & Yang, Y. J., (2013). Process and utility water requirements
for cellulosic ethanol production processes via fermentation pathway. Environ. Prog.
Sustain. Energy, 32(2), 396–405.
25. Singh, O. V., & Harvey, S. P., (2010). Sustainable biotechnology: Sources of renewable
energy. Sustain. Biotechnol. Sources Renew. Energy, pp. 1–323.
26. Putro, J. N., Soetaredjo, F. E., Lin, S. Y., Ju, Y. H., & Ismadji, S., (2016). Pretreatment
and conversion of lignocellulose biomass into valuable chemicals. RSC Adv., 6(52),
46834–46852.
27. Welker, C. M., Balasubramanian, V. K., Petti, C., Rai, K. M., De Bolt, S., & Mendu,
V., (2015). Engineering plant biomass lignin content and composition for biofuels and
bioproducts. Energies, 8(8), 7654–7676.
28. Shishir, P. S. C., Gregg T, B., Michael E, H., & Bruce E, D. (2011). Deconstruction
of Lignocellulosic Biomass to Fuels and Chemicals. Annual Review of Chemical and
Biomolecular Engineering, 2(1).
29. Kuila, A., & Sharma, V. (Eds.). (2017). Lignocellulosic Biomass Production and
Industrial Applications. John Wiley & Sons.
30. Kumar, S. J., Reetu, S., & Lakshmi, T., (2015). Lignocellulosic Agriculture Wastes
as Biomass Feedstocks for Second-generation Bioethanol Production: Concepts and
Apple Academic Press
Author Copy
34. Guerrero, A. B., Ballesteros, I., & Ballesteros, M., (2017). Optimal conditions of acid-
catalyzed steam explosion pretreatment of banana lignocellulosic biomass for fermentable
sugar production. J. Chem. Technol. Biotechnol., 92(9), 2351–2359.
35. Wu, M., Mintz, M., Wang, M., & Arora, S., (2009). Water consumption in the production
of ethanol and petroleum gasoline. Environ. Manage., 44(5), 981.
36. Neupane, B., Konda, N. V. S. N. M., Singh, S., Simmons, B. A., & Scown, C. D., (2017).
Life-Cycle Greenhouse Gas and Water Intensity of Cellulosic Biofuel Production Using
Cholinium Lysinate Ionic Liquid Pretreatment, 10176–10185.
37. Aden, A., Ruth, M., Ibsen, K., Jechura, J., Neeves, K., Sheehan, J. & Lukas, J. (2002).
Lignocellulosic biomass to ethanol process design and economics utilizing co-current
dilute acid prehydrolysis and enzymatic hydrolysis for corn stover (No. NREL/
TP-510–32438). National Renewable Energy Lab., Golden, CO. (US).
38. Dong, C., Wang, Y., Chan, K. L., Bhatia, A., & Leu, S. Y., (2018). Temperature profiling to
maximize energy yield with reduced water input in a lignocellulosic ethanol biorefinery.
Appl. Energy, 214, 63–72.
39. Singh, J., Suhag, M., & Dhaka, A., (2015). Augmented digestion of lignocellulose by
steam explosion, acid and alkaline pretreatment methods: A review. Carbohydr. Polym.,
117, 624–631.
40. Zhang, J., Wang, X., Chu, D., He, Y., & Bao, J., (2011). Dry pretreatment of lignocellulose
with extremely low steam and water usage for bioethanol production. Bioresour. Technol.,
102(6), 4480–4488.
41. Pan, S. Y., Lin, Y. J., Snyder, S. W., Ma, H. W., & Chiang, P. C., (2016). Assessing
the environmental impacts and water consumption of pretreatment and conditioning
processes of corn stover hydrolysate liquor in biorefineries. Energy, 116, 436–444.
42. Binod, P., Gnansounou, E., Sindhu, R., & Pandey, A., (2018). Enzymes for second-
generation biofuels: Recent developments and future perspectives. Bioresour. Technol.
Reports.
43. Sun, Y., & Cheng, J., (2002). Hydrolysis of lignocellulosic materials for ethanol
production: A review. Bioresour. Technol., 83(1), 1–11.
44. Yen, F. Y., Chen, H. H., Guo, G. L., Jang, M. F., & Chen, W. H., (2016). Method of
Pretreating Lignocellulose Using Solvent with Little Water Footprint (Vol. 1). United
States Pat. Appl. Publ.
45. Zhuang, X., et al., (2016). Liquid hot water pretreatment of lignocellulosic biomass for
bioethanol production accompanying with high valuable products. Bioresour. Technol.,
199, 68–75.
46. Tao, L., et al., (2011). Process and techno-economic analysis of leading pretreatment
technologies for lignocellulosic ethanol production using switchgrass. Bioresour.
Technol., 102(24), 11105–11114.
47. Cheah, W. Y., et al., (2020). Pretreatment methods for lignocellulosic biofuels production:
Apple Academic Press
Current advances, challenges and future prospects. Biofuel Res. J., 7(1), 1115–1127.
48. Cardona, E., Llano, B., Peñuela, M., Peña, J., & Rios, L. A., (2018). Liquid-hot-water
pretreatment of palm-oil residues for ethanol production: An economic approach to the
selection of the processing conditions. Energy, 160, 441–451.
49. Zimbardi, F., Viola, E., Nanna, F., Larocca, E., Cardinale, M., & Barisano, D., (2007).
Acid impregnation and steam explosion of corn stover in batch processes. Ind. Crops
Prod., 26(2), 195–206.
50. Solarte-Toro, J. C., Romero-García, J. M., Martínez-Patiño, J. C., Ruiz-Ramos, E.,
Castro-Galiano, E., & Cardona-Alzate, C. A., (2019). Acid pretreatment of lignocellulosic
Author Copy
biomass for energy vectors production: A review focused on operational conditions and
techno-economic assessment for bioethanol production. Renew. Sustain. Energy Rev.,
107, 587–601.
51. Abdul, R. A. Z., & Saidina, A. N. A., (2019). Environmental sustainability assessment
of palm lignocellulosic biomass pretreatment methods. Chem. Eng. Trans., 72, 13–18.
52. Omar, W. N. N. W., & Amin, N. A. S., (2016). Multi response optimization of oil palm
frond pretreatment by ozonolysis. Ind. Crops Prod., 85, 389–402.
53. Tan, H. T., Lee, K. T., & Mohamed, A. R., (2011). Pretreatment of lignocellulosic palm
biomass using a solvent-ionic liquid [BMIM] Cl for glucose recovery: An optimization
study using response surface methodology. Carbohydr. Polym., 83(4), 1862–1868.
54. Goshadrou, A., (2019). Bioethanol production from Cogon grass by sequential recycling
of black liquor and wastewater in a mild-alkali pretreatment. Fuel, 258, 116141.
55. Da Costa, S. L., et al., (2016). Next-generation ammonia pretreatment enhances
cellulosic biofuel production. Energy Environ. Sci., 9(4), 1215–1223.
56. Aden, A., et al., (2002a). Lignocellulosic Biomass to Ethanol Process Design and
Economics Utilizing Co-Current Dilute Acid Prehydrolysis and Enzymatic Hydrolysis
for Corn Stover. (No. NREL/TP-510–32438). National Renewable Energy Lab., Golden,
CO.(US).
57. Du, W., et al., (2011). The promoting effect of byproducts from Irpex lacteus on subse-
quent enzymatic hydrolysis of bio-pretreated cornstalks. Biotechnol. Biofuels, 4, 1–8.
58. Amriani, F., Salim, F. A., Iskandinata, I., Khumsupan, D., & Barta, Z., (2016). Physical
and biophysical pretreatment of water hyacinth biomass for cellulase enzyme production.
Chem. Biochem. Eng. Q., 30(2), 237–244.
59. Aden, A., & others, (2007). Water usage for current and future ethanol production.
Southwest Hydrol., 6(5), 22, 23.
60. Gu, Y. M., Kim, H., Sang, B. I., & Lee, J. H., (2018). Effects of water content on ball
milling pretreatment and the enzymatic digestibility of corn stover. Water-Energy Nexus,
1(1), 61–65.
Author Copy
LUIS FERNANDO AMADOR CASTRO and DANAY CARRILLO NIEVES
Tecnologico de Monterrey, Escuela de Ingenieria y Ciencias,
Av. General Ramon Corona No. 2514, Zapopan–45201, Jal., Mexico
E-mail: danay.carrillo@tec.mx (Danay Carrillo Nieves)
ABSTRACT
15.1 INTRODUCTION
Author Copy
use of biomass has been proposed as one of the alternatives to fossil fuels
being lignocellulose of particular importance as it is known to be the most
abundant biomass on Earth [6].
Lignocellulose is an organic material that constitutes the plant cell walls.
It can be readily obtained from feedstocks and forestry as well as from their
wastes [7]. Biofuels that can be obtained from this material include ethanol,
butanol, and biodiesel; among the previous ethanol has the highest relevance
[8]. Ethanol can be readily mixed with gasoline reducing hydrocarbon and
carbon monoxide emissions [9]. Ethanol-gasoline blends can also contribute
to decreasing the effects of climate change given that feedstocks used for
ethanol production sequester CO2 for its growth [10]. Additionally, the use
of lignocellulosic wastes and non-edible crops for ethanol production can
favor rural economic development and at the same time does not comprise
food security [11]. However, the implementation of this technology faces
significant challenges.
Serving as a defense mechanism against pathogens, lignocellulose is
highly recalcitrant, and its hydrolysis requires the employment of chemical
or enzymatic procedures. Furthermore, to improve the ethanol yield
biomass should be pretreated either by physical or chemical means [12].
The requirement of a multiple-step process for ethanol production from
LCB makes the process not economically competitive when compared
to currently existing fuels [11]. Aiming towards cost reduction, multiple
strategies have been proposed, including the elimination of the pretreat-
ment procedures, the increase in cellulose hydrolysis yield by optimizing
the enzymes involved in the process and improving the ethanol yields [13].
However, one of the most promising strategies is to perform the conver-
sion of lignocellulose into ethanol by implementing a single-step process
known as CBP [14].
provide the reader with some insights regarding the structure of the plant cell
walls and the enzymes that are involved in their synthesis and degradation.
Further, we explain the concepts, advantages, and challenges around CBP;
and finally, we review some fungi that have been used or have the potential
to be used in this type of process as they are known to be the most efficient
lignocellulosic enzymes producers.
Author Copy
15.2 LIGNOCELLULOSIC MATERIALS, THE STRUCTURE, AND
RECALCITRANCE OF CELL WALL
The cell wall is a critical component of plant cells. Apart from encasing cell’s
protoplasm, it participates in biological processes such as differentiation,
intracellular communication, water, and nutrient transport and as a defense
mechanism against pathogens [15]. Cell wall composition varies greatly
among different plant species and within cells from the same plant depending
on the physiological role they exert. Furthermore, there are some plant cells
that contain two cell walls, primary, and secondary, which composition is
also different [15]. Plant cell walls have a complex structure consisting of a
mixture of carbohydrates, proteins, and aromatic compounds. Carbohydrates
that form the plant cell walls include cellulose, hemicellulose, and pectin and
can account for as much as 90% of the content of the primary cell wall [16].
Aromatic compounds like lignin which are present in the walls of many plant
cells provide plants with further structural support. Proteins present on the
cell walls can serve as structural molecules or signaling receptors which are
essential for plant development. This section will briefly describe the main
components of the plant cell walls.
15.2.1 LIGNIN
guaiacyl (G), and syringyl (S) lignin subunits, respectively derive [18].
Recent evidence indicates that apart from hydroxycinnamyl alcohols, lignin
can be synthesized to a less extent from other phenolic monomers, further
increasing the complexity of this polymer [19]. Still, all currently known
lignin building blocks derived from the general phenylpropanoid pathway
in which phenylalanine serves as the substrate for this pathway in plants and
tyrosine as an additional substrate in grasses. The amount of lignin as well as
Author Copy
its composition vary significantly among and within plant species cell types
and even cell walls; being influenced by diverse environmental factors [19].
Generally, lignin polymers from dicotyledonous angiosperms are composed
mostly of G and S subunits having only traces of H subunits. Gymnosperm
lignins lack S subunits and are composed as much as 90% from G subunits.
Lignins from grasses contain G and S subunits at similar levels and a lower
amount of H [20]. In lignin, these subunits are typically connected by
carbon-oxygen ether linkages and carbon-carbon linkages being β-O-4 is
the most common linkage. Other molecular bonds include α-O-4, β-5, 5-5,
4-O-5, β-1, and β-β linkages [20]. The interest in the study of lignin arises as
this polymer, which accounts for up to 30% of LCB, can greatly hinder the
process of the bioconversion of LCB into ethanol. Therefore, reducing its
content or changing its composition can represent significant improvements
regarding the use of lignocellulosic materials in multiple industries.
15.2.2 CELLULOSE
Cellulose is the most abundant polymer of the plant cell being present in both
primary and secondary walls. It provides structural support allowing plants
to withstand physical stress and act as a barrier against other organisms.
Cellulose is not exclusive to plants as certain bacteria, fungi, and animals
can synthesize it [21]. This polymer consists of chains of repeating D-gluco-
pyranose molecules which are linked by β-1,4-glycosidic bonds. Chains of
cellulose comprise a non-reducing end, which is a glucose molecule with
its original C4-OH group and a reducing end where C1-OH is present; the
rest of the polymer contains anhydroglucose molecules containing three
hydroxyl groups. Each anhydroglucose molecule is rotated 180° in the plane,
being two adjacent molecules, the cellulose monomer called cellobiose [22].
Author Copy
chains have a lower molecular weight and less DP when compared to those
of the secondary cell wall [21]. Being the main polymer present in LCB,
cellulose regularly accounts for 30–50% of its dry weight. Further, glucose
molecules proceeding from cellulose degradation can undergo fermentation,
thus being an important substrate for biofuel production.
15.2.3 HEMICELLULOSE
15.2.4 PECTIN
Author Copy
constituting about 35% of primary walls in non-graminaceous monocots
and dicots, 2–10% of grass primary walls, and as much as 5% of woody
tissue walls. As previous cell wall constituents, pectin is involved in vital
plant functions including growth processes, morphology, development,
and defense against environment and pathogens. In the cell walls, it also
provides strength and flexibility [28]. Pectin polysaccharides include
homogalacturonan, rhamnogalacturonan I (RGI), rhamnogalacturonan
II (RGII), and xylogalacturonan (XGA). The total pectin content and the
relative proportion of the polysaccharides that constitute it varies depending
on environmental factors, from plant to plant and within same the plant.
Homogalacturonan is the most common polysaccharide, it can account for
up to 65% of pectin; RGI constitutes from 20% to 35%. XGA and RGII
are present in minor amounts, each accounting for less than 10% of pectin
[27]. All previous polysaccharides contain galacturonic acid linked at the
O-1 and the O-4 position. Homogalacturonan is a linear polymer consisting
only of α-1,4-linked galacturonic acid molecules. The common length of the
chain is of about 100 galacturonic acid molecules, however, shorter chains
can be found. RGII structure is far more complex, it consists of a backbone
of homogalacturonan that contains branches formed from different sugars.
XGA is composed of a homogalacturonan chain substituted at O-3 with a
β-linked xylose. Finally, RGI consists of a backbone of the disaccharide
repeats of rhamnose and galacturonic acid, that can be substituted with side
chains containing L-galactose and L-arabinose residues [27].
Apart from the cited polysaccharides and lignin, plant cell walls contain other
minor components that may be taken into consideration when designing a
process where LCB is the raw material. Plants contain a variety of minerals,
sometimes referred to as ashes, which are important for their development.
The presence of high contents of ashes can have negative effects on enzymatic
hydrolysis and fermentation processes [7]. Plants also contain pigments such as
Apple Academic Press
flavonoids and anthocyanins that, when extracted, can affect the bioconversion
of lignocellulose into certain products, therefore pretreatment processes should
be considered depending on the final use of the biomass. A variety of proteins
can also be found on the plant cell wall, these proteins exert mainly structural
functions, although some are involved in morphogenesis having signaling
functions. Most cell wall proteins are glycosylated and contain hydroxyproline
except for glycine-rich proteins [29]. The presence of proteins adds to further
Author Copy
increase the complexity of the plant cell walls, which degradation will involve
the participation of multiple enzymes as we well address in further section.
The recalcitrance of the cell wall is one of the first challenges for the
bioprocessing of lignocellulosic materials for their use in biofuel production.
As aforementioned, plant cell wall is composed mainly of lignin, cellulose,
hemicellulose, and pectin; therefore, its complete enzymatic degradation
entails the participation of several lignocellulosic enzymes. However, nature
can give us the solution, as lignocellulosic microorganisms like fungi possess
different enzymes that can make feasible the degradation of this structure.
Lignin recalcitrance to degradation is due to its complex structure and linkage
heterogeneity [30]. Lignin-degrading enzymes can be classified mainly into
two groups namely heme peroxidases and laccases [31, 32]. Ligninolytic heme
peroxidases comprise lignin, manganese, and versatile and dye decolorizing
peroxidases (DyP) [31]. Laccases assist lignin degradation and are part of
the multicopper oxidase superfamily [30]. Additional enzymes that produce
hydrogen peroxide (H2O2), which is required for peroxidase-based lignin
degradation, include glyoxal oxidase, pyranose-2 oxidase, and aryl alcohol
oxidase [33]. Cellulose and hemicellulose contain most of the sugars of the
plant cell walls, but to make these sugars available for fermentation, the
complex structures of these polymers need to be degraded. The hydrolysis
of these cell walls’ components requires the synergistic action of multiple
cellulases and hemicellulases. Pectin degradation is performed by a set of
pectinases, enzymes that have found applications in different industries.
Author Copy
Peroxidases are widely distributed among plants, animals, and microorgan-
isms. These enzymes catalyze the oxidation of different substrates, using
H2O2 or other peroxides as electron acceptors [35]. Most of the peroxidases
are heme enzymes that contain an iron protoporphyrin IX prosthetic group,
although there are many nonheme peroxidases [36]. Heme peroxidases can
be mainly classified into two superfamilies animal and non-animal (plant)
peroxidases [37, 38]. Non-animal peroxidases are further subdivided into
three classes. Class I are intracellular peroxidases that have been found in
bacteria, fungi, plants, and some protists, examples include cytochrome c,
catalase, and ascorbate peroxidases. Class II refers to extracellular fungal
peroxidases in which lignin, manganese, and versatile peroxidases (VPs) are
included. Class III are secreted peroxidases only found in plants [39]. Animal
peroxidases have also been subclassified as in the case of plant peroxidases,
however, its subclassification is more complex [37]. Some heme peroxidases
such as DyPs do not fit within the previously mentioned superfamily clas-
sification and constitute its family [40]. Different peroxidase phylogenetic
classifications have also been proposed [41, 42].
Author Copy
[Fe (IV)] using H2O2 as the electron acceptor. In the second step, the interme-
diate, known as compound I, is reduced by a substrate which translocases one
electron to it to form another intermediate referred to as compound II. The
final step compound II receives another electron from the reduced substrate
and consequently, the enzyme returns to its original oxidation state [50].
Author Copy
of about 4.5 [60]. This type of peroxidases can oxidize substrates in a broad
range of potentials, including low and high redox potentials compounds. The
catalytic cycle of VP is like that of previously described lignin and manga-
nese heme peroxidases. It involves the first two-electron oxidation followed
by two reduction reactions in which intermediate compounds are formed
to return the enzyme to its native oxidation state. The main difference is
that these enzymes have the versatility of oxidizing either Mn2+ or other
substrates to carry out their reduction reactions [59].
15.3.2 LACCASES
Author Copy
are only extracellular, laccases can be localized extra or intracellularly. The
localization of these enzymes may be related to the physiological function
they exert, and this localization also determines the variety of substrates
that can be used by them [66]. Generally, fungal laccases have masses
around 60–70 kDa, an isoelectric point about 4, and an acidic optimum pH
between 2.5–4 [66].
The catalytic cycle of laccases reduces one molecule of O2 to two
molecules of water with the concomitant oxidation of four molecules of a
substrate to produce four radicals. This redox process is aided by a cluster
of four copper atoms that serve as the catalytic core of the enzyme [67].
Laccases can catalyze direct oxidation of phenolic compounds and certain
amines [68]. Nevertheless, oxidization of high redox potential compounds
such as non-phenolic model compounds and β-1 lignin dimers requires the
presence of a mediator. Some commonly used mediators are 3-hydroxyan-
thranilic acid (HAA), N-hydroxybenzotriazole, and 2,20-azino-bis-(3-ethyl-
benzothiazoline-6-sulphonic acid) (ABTS) [67]. However, the use of these
mediators increases the overall cost of the process.
including textile, wine, and brewery, food, pulp, and paper, agriculture, and
biofuel industries [70, 71].
Apple Academic Press
15.3.3.1 CELLULASES
Author Copy
some engineered strains have in producing these enzymes [73]. In bacteria,
cellulases are present as extracellular aggregated structures attached to the
cells called cellulosomes [74]. Fungal cellulases have a relatively simple
molecular structure containing a catalytic domain and a cellulose-binding
domain. Extensive reviews describing the structure and characteristics of
these enzymes are available on refs [75, 76]. The catalytic domain is linked
to the cellulose-binding domain by a linker peptide. The anchoring of the
cellulose-binding domain to cellulose permits the enzyme to optimally
perform its catalytic function [75]. Cellulases degrade cellulose using
different synergistic approaches. Endoglucanases, also called non-processive
cellulases, can cleave the internal O-glycosidic bonds producing polysac-
charide chains of variable lengths. Exoglucanases or processive cellulases
act by coupling to the cellulose chain ends and hydrolyzing cellulose into
cellobiose. β-glucosidases can then cleave cellobiose into glucose that can
be used by the microorganism [76].
15.3.3.2 HEMICELLULASES
acetyl xylan esterases (EC 3.1.1.72) [78, 79]. Many of these hemicellulolytic
enzymes have been found on various types of microorganisms, including
fungi and bacteria [78]. Describing the catalytic mechanisms and structure
of the previously mentioned enzymes is beyond the scope of this chapter.
Apple Academic Press
15.3.4 PECTINASES
Author Copy
bility, porosity, and electrostatic potential [80]. Because of its contribution to
porosity and charge, pectin has been a target for studies aiming at cell wall
degradation [81]. Pectinases are a group of enzymes capable of degrading
pectin that has found application in the fruit, paper, and pulp, textile, and
biofuel industries [81, 82]. Pectinases are natively produced by a variety
of organisms including fungi, bacteria, insects, protozoa, and plants [83].
However, for the industrial production of these enzymes, fungi, and bacteria
are the most commonly employed organisms [84]. Different classifications
for pectinases have been proposed that are based on their preferred substrate,
their mode of action, and cleavage characteristics [85]. Microorganisms can
produce different forms of pectinases that have been encountered to work
in a wide range of pH and temperature values [82]. Optimum pH values
may vary between 4–10 and optimal temperature from 30 to 70°C [82].
Such variability can be related to the natural environmental factors to which
the microorganism is exposed [83]. Because of the previous recombinant
pectinases are produced from different sources so that they can be applied in
processes that involve harsh temperature and pH conditions [84].
phenylalanine and sometimes tyrosine the substrates for this pathway [19].
Furthermore, laccases, and peroxidases, enzymes that are involved in the
lignin degradation process, are also involved in lignin biosynthesis by
catalyzing the oxidation of phenolic monolignols into phenolic monomers,
a process essential for the initiation of in vivo polymerization of lignin [30].
The study regarding cell wall synthesis is of great interest due as one
strategy to overcome its recalcitrance may be to develop feedstock varieties
Author Copy
with more labile cell walls that apart from maintaining is agricultural yields
can be useful for biofuel production [89]. Some of the enzymes that partici-
pate in the degradation of the main components of the plant cell wall have
also been identified to be involved in its biosynthesis. This is the case of the
membrane-anchored endoglucanase Korrigan which appears to be related to
cellulose biosynthesis in the primary and secondary cell walls of Arabidopsis
thaliana. Mutant genotypes have exhibited defects in cell wall formation,
including reduced cellulose content and increased pectin synthesis deriving
in abnormal plant morphology [90, 91]. Comprehensive reviews can provide
the reader with a complete picture of the current understanding of the biosyn-
thesis of the main components of the cell wall [19, 25, 80, 92].
Author Copy
make able to develop a consolidated bioprocess to simultaneously perform
the saccharification and fermentation processes.
being the cost of the enzyme the most significant [101]. All the previous
has led to the development of CBP, where enzyme production along with
saccharification and fermentation are produced simultaneously within the
same vessel [101]. For CBP to be feasible, it is required that either multiple
Apple Academic Press
Author Copy
growth of its producing microorganisms once a certain concentration has
been reached; thus, discontinuing the fermentation process which may result
in low ethanol yield [99]. These factors have led to the quest of finding micro-
organisms that are optimal for this type of bioprocess. However, microbes
capable of overcoming previous difficulties and that possess the enzymatic
machinery to perform both saccharification and fermentation efficiently have
not yet been encountered in nature [101, 102]. Therefore, different strategies
have been proposed to develop these ideal microorganisms. Mainly there is a
focus on two strategies, the native strategy, and the recombinant strategy. The
first one aims to genetically engineer a microorganism natively capable of
producing lignocellulosic enzymes to make it ethanologenic. The second is
targeted to engineer an ethanol-producing microorganism to make it capable
of degrading lignocellulose [97, 102].
Following the native or recombinant strategies has some challenges. By
engineering microorganisms according to the native strategy, scientists need
to use the currently available genetic engineering tools to produce a strain
capable of yielding high ethanol concentrations and that is robust enough for
its industrial application [102]. Most promising candidate microorganisms
for following this strategy are lignocellulosic fungi and bacteria, as microor-
ganisms like bacterium Clostridium thermocellum and a variety of white-rot
fungi have been already used in ethanol production via CBP [103, 104]. On
the other hand, the production of organisms by the recombinant strategy
faces the challenge that the selected organisms need to produce enough
quantities of a wide variety of enzymes to degrade the main components of
the plant cell walls [97]. The recombinant strategy has focused mainly on
modifying yeasts and bacteria, with most of the advances being performed
in S. cerevisiae [102, 105].
Besides previously mentioned challenges, other aspects limit the full
potential of CBP. Genetically engineered microorganisms may present
Author Copy
to be engineered to overcome these difficulties. In the following section, we
will present some fungi that are good candidates or that already have been
employed in the CBP of ethanol production from LCB.
At the hand of the abundant presence of biomass and the pressing need for
different fuel sources, the finding of fungi with the potential to produce
lignocellulolytic enzymes is a viable solution, regarding the problem. The
fungi are grouped in relation to the type and characteristics of produced
rot, in spite of the classification system is not accurately aimed. Table 15.1
depicts the following groups of fungi-rot.
White-rot fungi produce a multi enzymatic system allowing for break down
the three principal polymers of the natural lignocellulosic substrates: lignin,
Apple Academic Press
Author Copy
Ascomycetes and Deuteromycetes belong to soft-rot fungi and they are the
noblest fungi in terms of the level of decomposition they cause. Even, they
emerge before the rotting process carried out by white and brown fungi. Despite
the low degradation, soft-rot fungi degrading wood in severe environmental
conditions. These kinds of fungi require fix nitrogen to synthesize cellulase
enzymes, in charge of depolymerizing the cellulose of the secondary cell
walls, making huge cavities [108, 113, 114].
Numerous fungi cause white rot compared to brown rot. However, they
play an important role in the degradation of deadwood, especially, in
coniferous ecosystems. The brown rot fungi firstly attack the lignin through
demethoxylation, occurring a lignin partial modification, without produce
lignin-degrading enzymes. Posteriorly, the cellulose tackle is carried out by
the transformation of methanol to peroxide with a methanol oxidase over-
expression via Fenton chemistry. Thanks to the accelerated fractionation of
cellulose, the wood loses strength rapidly [95, 114].
KEYWORDS
• consolidated bioprocessing
Apple Academic Press
• hydrogen peroxide
• lignin peroxidase
• manganese peroxidase
• rhamnogalacturonan I
• separate hydrolysis and fermentation
Author Copy
REFERENCES
12. Canilha, L., Chandel, A. K., Suzane, D. S. M. T., Antunes, F. A. F., Luiz Da, C. F. W., Das,
G. A. F. M., & Da Silva, S. S., (2012). Bioconversion of sugarcane biomass into ethanol:
An overview about composition, pretreatment methods, detoxification of hydrolysates,
enzymatic saccharification, and ethanol fermentation. J. Biomed. Biotechnol., 2012,
Apple Academic Press
1–15. https://doi.org/10.1155/2012/989572.
13. Xu, Q., Singh, A., & Himmel, M. E., (2009). Perspectives and new directions for the
production of bioethanol using consolidated bioprocessing of lignocellulose. Curr.
Opin. Biotechnol., 20(3), 364–371. https://doi.org/10.1016/j.copbio.2009.05.006.
14. Salvachúa, D., Karp, E. M., Nimlos, C. T., Vardon, D. R., & Beckham, G. T., (2015).
Towards lignin consolidated bioprocessing: Simultaneous lignin depolymerization and
product generation by bacteria. Green Chem., 17(11), 4951–4967. https://doi.org/ 10.1039/
C5GC01165E.
15. Cosgrove, D. J., (2005). Growth of the plant cell wall. Nat. Rev. Mol. Cell Biol., 6(11),
Author Copy
850–861. https://doi.org/10.1038/nrm1746.
16. Caffall, K. H., & Mohnen, D., (2009). The structure, function, and biosynthesis of plant
cell wall pectic polysaccharides. Carbohydr. Res., 344(14), 1879–1900. https://doi.
org/10.1016/j.carres.2009.05.021.
17. Vanholme, R., Morreel, K., Ralph, J., & Boerjan, W., (2008). Lignin engineering. Curr.
Opin. Plant Biol., 11(3), 278–285. https://doi.org/10.1016/j.pbi.2008.03.005.
18. Ragauskas, A. J., Beckham, G. T., Biddy, M. J., Chandra, R., Chen, F., Davis, M. F., Davison,
B. H., et al., (2014). Lignin valorization: Improving lignin processing in the biorefinery.
Science, 344(6185), 1246843–1246843. https://doi.org/10.1126/science.1246843.
19. Vanholme, R., De Meester, B., Ralph, J., & Boerjan, W., (2019). Lignin biosynthesis
and its integration into metabolism. Curr. Opin. Biotechnol., 56, 230–239. https://doi.
org/10.1016/j.copbio.2019.02.018.
20. Boerjan, W., Ralph, J., & Baucher, M., (2003). Lignin biosynthesis. Annu. Rev. Plant
Biol., 54(1), 519–546. https://doi.org/10.1146/annurev.arplant.54.031902.134938.
21. Brett, C. T., (2000). Cellulose microfibrils in plants: Biosynthesis, deposition, and
integration into the cell wall. In: International Review of Cytology (Vol. 199, pp.
161–199). Elsevier. https://doi.org/10.1016/S0074-7696(00)99004-1.
22. Klemm, D., Heublein, B., Fink, H. P., & Bohn, A., (2005). Cellulose: Fascinating
biopolymer and sustainable raw material. Angew. Chem. Int. Ed., 44(22), 3358–3393.
https://doi.org/10.1002/anie.200460587.
23. Foyle, T., Jennings, L., & Mulcahy, P., (2007). Compositional analysis of lignocellulosic
materials: Evaluation of methods used for sugar analysis of waste paper and straw.
Bioresour. Technol., 98(16), 3026–3036. https://doi.org/10.1016/j.biortech.2006.10.013.
24. Somerville, C., (2006). Cellulose synthesis in higher plants. Annu. Rev. Cell Dev. Biol.,
22(1), 53–78. https://doi.org/10.1146/annurev.cellbio.22.022206.160206.
25. Pauly, M., Gille, S., Liu, L., Mansoori, N., De Souza, A., Schultink, A., & Xiong, G.,
(2013). Hemicellulose biosynthesis. Planta, 238(4), 627–642. https://doi.org/10.1007/
s00425-013-1921-1.
26. Scheller, H. V., & Ulvskov, P., (2010). Hemicelluloses. Annu. Rev. Plant Biol., 61(1),
263–289. https://doi.org/10.1146/annurev-arplant-042809-112315.
27. Mohnen, D., (2008). Pectin structure and biosynthesis. Curr. Opin. Plant Biol., 11(3),
266–277. https://doi.org/10.1016/j.pbi.2008.03.006.
28. Harholt, J., Suttangkakul, A., & Scheller, H. V., (2010). Biosynthesis of pectin. Plant
Physiol., 153(2), 384–395. https://doi.org/10.1104/pp.110.156588.
29. Cassab, G. I., (1998). Plant cell wall proteins. Annu. Rev. Plant Physiol. Plant Mol.
Biol., 49(1), 281–309. https://doi.org/10.1146/annurev.arplant.49.1.281.
30. De Gonzalo, G., Colpa, D. I., Habib, M. H. M., & Fraaije, M. W., (2016). Bacterial
enzymes involved in lignin degradation. J. Biotechnol., 236, 110–119. https://doi.org/
Apple Academic Press
10.1016/j.jbiotec.2016.08.011.
31. Abdel-Hamid, A. M., Solbiati, J. O., & Cann, I. K. O., (2013). Insights into lignin
degradation and its potential industrial applications. In: Sariaslani, S., & Gadd, G. M.,
(eds.), Advances in Applied Microbiology (Vol. 82, pp. 1–28). Academic Press. https://
doi.org/10.1016/B978-0-12-407679-2.00001-6.
32. Kinnunen, A., Maijala, P., Jarvinen, P., & Hatakka, A., (2017). Improved efficiency
in screening for lignin-modifying peroxidases and laccases of basidiomycetes. Curr.
Biotechnol., 6(2), 105–115. https://doi.org/10.2174/2211550105666160330205138.
33. Tuomela, M., & Hatakka, A., (2019). Oxidative fungal enzymes for bioremediation.
Author Copy
In: Moo-Young, M., (ed.), Comprehensive Biotechnology (3rd edn., pp. 224–239);
Pergamon: Oxford. https://doi.org/10.1016/B978-0-444-64046-8.00349-9.
34. Shen, H., Poovaiah, C. R., Ziebell, A., Tschaplinski, T. J., Pattathil, S., Gjersing, E.,
Engle, N. L., et al., (2013). Enhanced characteristics of genetically modified switchgrass
(Panicum Virgatum L.) for high biofuel production. Biotechnol. Biofuels, 6(1), 71.
https://doi.org/10.1186/1754-6834-6-71.
35. Hamid, M., & Khalil-Ur-Rehman, (2009). Potential applications of peroxidases. Food
Chem., 115(4), 1177–1186. https://doi.org/10.1016/j.foodchem.2009.02.035.
36. Hofrichter, M., Ullrich, R., Pecyna, M. J., Liers, C., & Lundell, T., (2010). New and
classic families of secreted fungal heme peroxidases. Appl. Microbiol. Biotechnol., 87(3),
871–897. https://doi.org/10.1007/s00253-010-2633-0.
37. Passardi, F., Theiler, G., Zamocky, M., Cosio, C., Rouhier, N., Teixera, F., Margis-
Pinheiro, M., et al., (2007). PeroxiBase: The peroxidase database. Phytochemistry,
68(12), 1605–1611. https://doi.org/10.1016/j.phytochem.2007.04.005.
38. Savelli, B., Li, Q., Webber, M., Jemmat, A. M., Robitaille, A., Zamocky, M., Mathé, C.,
& Dunand, C., (2019). RedoxiBase: A database for ROS homeostasis regulated proteins.
Redox Biol., 26, 101247. https://doi.org/10.1016/j.redox.2019.101247.
39. Passardi, F., Bakalovic, N., Teixeira, F. K., Margis-Pinheiro, M., Penel, C., & Dunand,
C., (2007). Prokaryotic origins of the non-animal peroxidase superfamily and organelle-
mediated transmission to eukaryotes. Genomics, 89(5), 567–579. https://doi.org/10.1016/j.
ygeno.2007.01.006.
40. Sugano, Y., (2009). DyP-type peroxidases comprise a novel heme peroxidase family.
Cell. Mol. Life Sci., 66(8), 1387–1403. https://doi.org/10.1007/s00018-008-8651-8.
41. Zámocký, M., & Obinger, C., (2010). Molecular phylogeny of heme peroxidases. In:
Torres, E., & Ayala, M., (eds.), Biocatalysis Based on Heme Peroxidases: Peroxidases
as Potential Industrial Biocatalysts (pp. 7–35). Springer: Berlin, Heidelberg. https://doi.
org/10.1007/978-3-642-12627-7_2.
42. Zámocký, M., Hofbauer, S., Schaffner, I., Gasselhuber, B., Nicolussi, A., Soudi,
M., Pirker, K. F., et al., (2015). independent evolution of four heme peroxidase
superfamilies. Arch. Biochem. Biophys., 574, 108–119. https://doi.org/10.1016/j.abb.
2014.12.025.
43. Paliwal, R., Rawat, A. P., Rawat, M., & Rai, J. P. N., (2012). Bioligninolysis: Recent
updates for biotechnological solution. Appl. Biochem. Biotechnol., 167(7), 1865–1889.
https://doi.org/10.1007/s12010-012-9735-3.
44. Johansson, T., & Nyman, P. O., (1993). Isozymes of lignin peroxidase and manganese
(II) peroxidase from the white-rot basidiomycete Trametes versicolor. Arch. Biochem.
Biophys., 300(1), 49–56. https://doi.org/10.1006/abbi.1993.1007.
45. Tien, M., & Kirk, T. K., (1983). Lignin-degrading enzyme from the hymenomycete
Apple Academic Press
Author Copy
Mancinelli, S., (2001). Isotope-effect profiles in the oxidative N-demethylation of
N,N-dimethylanilines catalyzed by lignin peroxidase and a chemical model. Eur. J.
Org. Chem., 2001(12), 2305–2310. https://doi.org/10.1002/1099-0690(200106)2001:
12<2305::AID-EJOC2305>3.0.CO;2-E.
49. Wong, D. W. S., (1995). Food Enzymes. Springer US: Boston, MA. https://doi.
org/10.1007/978-1-4757-2349-6.
50. Falade, A. O., Nwodo, U. U., Iweriebor, B. C., Green, E., Mabinya, L. V., & Okoh, A.
I., (2017). Lignin peroxidase functionalities and prospective applications. Microbiology
Open, 6(1), e00394. https://doi.org/10.1002/mbo3.394.
51. Dashtban, M., Schraft, H., Syed, T. A., & Qin, W., (2010). Fungal biodegradation and
enzymatic modification of lignin. Int. J. Biochem. Mol. Biol., 1(1), 36–50.
52. Zhang, H., Zhang, J., Zhang, X., & Geng, A., (2018). Purification and characterization
of a novel manganese peroxidase from white-rot fungus Cerrena unicolor BBP6 and its
application in dye decolorization and denim bleaching. Process Biochem., 66, 222–229.
https://doi.org/10.1016/j.procbio.2017.12.011.
53. Lobos, S., Larraín, J., Salas, L., Cullen, D., & Vicuña, R., (1994). Isoenzymes of manga-
nese-dependent peroxidase and laccase produced by the lignin-degrading basidiomycete
Ceriporiopsis subvermispora. Microbiol. Read. Engl., 140(Pt 10), 2691–2698. https://doi.
org/10.1099/00221287-140-10-2691.
54. Chowdhary, P., Shukla, G., Raj, G., Ferreira, L. F. R., & Bharagava, R. N., (2018).
Microbial manganese peroxidase: A ligninolytic enzyme and its ample opportunities in
research. SN Appl. Sci., 1(1), 45. https://doi.org/10.1007/s42452-018-0046-3.
55. Reddy, G. V. B., Sridhar, M., & Gold, M. H., (2003). Cleavage of nonphenolic β-1 diarylpro-
pane lignin model dimers by manganese peroxidase from Phanerochaete chrysosporium.
Eur. J. Biochem., 270(2), 284–292. https://doi.org/10.1046/j.1432-1033. 2003.03386.x.
56. Moreira, P. R., Duez, C., Dehareng, D., Antunes, A., Almeida-Vara, E., Frère, J. M.,
Malcata, F. X., & Duarte, J. C., (2005). Molecular characterization of a versatile
peroxidase from a Bjerkandera strain. J. Biotechnol., 118(4), 339–352. https://doi.org/
10.1016/j.jbiotec.2005.05.014.
57. Kamitsuji, H., Honda, Y., Watanabe, T., & Kuwahara, M., (2005). Mn2+ is dispensable
for the production of active MnP2 by Pleurotus ostreatus. Biochem. Biophys. Res.
Commun., 327(3), 871–876. https://doi.org/10.1016/j.bbrc.2004.12.084.
58. Ruiz-Dueñas, F. J., Martínez, M. J., & Martínez, A. T., (1999). Molecular characterization
of a novel peroxidase isolated from the ligninolytic fungus Pleurotus eryngii. Mol.
Microbiol., 31(1), 223–235. https://doi.org/10.1046/j.1365-2958.1999.01164.x.
59. Wong, D. W. S., (2009). Structure and action mechanism of ligninolytic enzymes. Appl.
Biochem. Biotechnol., 157(2), 174–209. https://doi.org/10.1007/s12010-008-8279-z.
60. Verdín, J., Pogni, R., Baeza, A., Baratto, M. C., Basosi, R., & Vázquez-Duhalt, R.,
(2006). Mechanism of versatile peroxidase inactivation by Ca2+ depletion. Biophys.
Apple Academic Press
Author Copy
application in enzymatic hydrolysis of wheat straw. Appl. Environ. Microbiol., 79(14),
4316–4324. https://doi.org/10.1128/AEM.00699-13.
64. Liers, C., Bobeth, C., Pecyna, M., Ullrich, R., & Hofrichter, M., (2010). DyP-like
peroxidases of the jelly fungus Auricularia Auricula-Judae oxidize nonphenolic lignin
model compounds and high-redox potential dyes. Appl. Microbiol. Biotechnol., 85(6),
1869–1879. https://doi.org/10.1007/s00253-009-2173-7.
65. Claus, H., (2004). Laccases: Structure, reactions, distribution. Micron, 35(1), 93–96.
https://doi.org/10.1016/j.micron.2003.10.029.
66. Baldrian, P., (2006). Fungal laccases - occurrence and properties. FEMS Microbiol.
Rev., 30(2), 215–242. https://doi.org/10.1111/j.1574-4976.2005.00010.x.
67. Riva, S., (2006). Laccases: Blue enzymes for green chemistry. Trends Biotechnol., 24(5),
219–226. https://doi.org/10.1016/j.tibtech.2006.03.006.
68. Giardina, P., Faraco, V., Pezzella, C., Piscitelli, A., Vanhulle, S., & Sannia, G., (2010).
Laccases: A never-ending story. Cell. Mol. Life Sci., 67(3), 369–385. https://doi.org/
10.1007/s00018-009-0169-1.
69. Teeri, T. T., (1997). Crystalline cellulose degradation: New insight into the function
of cellobiohydrolases. Trends Biotechnol., 15(5), 160–167. https://doi.org/10.1016/
S0167-7799(97)01032-9.
70. Li, X., Chang, S. H., & Liu, R., (2018). Industrial applications of cellulases and
hemicellulases. In: Fang, X., & Qu, Y., (eds.), Fungal Cellulolytic Enzymes (pp. 267–282).
Springer Singapore: Singapore. https://doi.org/10.1007/978-981-13-0749–2_15.
71. Bhat, M. K., (2000). Cellulases and related enzymes in biotechnology. Biotechnol. Adv.,
18(5), 355–383. https://doi.org/10.1016/S0734-9750(00)00041-0.
72. Kuhad, R. C., Gupta, R., & Singh, A., (2011). Microbial cellulases and their industrial
applications. Enzyme Res., 2011, 1–10. https://doi.org/10.4061/2011/280696.
73. Wilson, D. B., (2009). Cellulases and biofuels. Curr. Opin. Biotechnol., 20(3), 295–299.
https://doi.org/10.1016/j.copbio.2009.05.007.
74. Doi, R. H., & Kosugi, A., (2004). Cellulosomes: Plant-cell-wall-degrading enzyme
complexes. Nat. Rev. Microbiol., 2(7), 541–551. https://doi.org/10.1038/nrmicro925.
75. Juturu, V., & Wu, J. C., (2014). Microbial cellulases: Engineering, production and
applications. Renew. Sustain. Energy Rev., 33, 188–203. https://doi.org/10.1016/j.rser.
2014.01.077.
76. Payne, C. M., Knott, B. C., Mayes, H. B., Hansson, H., Himmel, M. E., Sandgren,
M., Ståhlberg, J., & Beckham, G. T., (2015). Fungal cellulases. Chem. Rev., 115(3),
1308–1448. https://doi.org/10.1021/cr500351c.
77. Gao, D., Uppugundla, N., Chundawat, S. P., Yu, X., Hermanson, S., Gowda, K., Brumm,
P., et al., (2011). Hemicellulases and auxiliary enzymes for improved conversion of
lignocellulosic biomass to monosaccharides. Biotechnol. Biofuels, 4(1), 5. https://doi.
org/10.1186/1754-6834-4-5.
Apple Academic Press
78. Shallom, D., & Shoham, Y., (2003). Microbial hemicellulases. Curr. Opin. Microbiol.,
6(3), 219–228. https://doi.org/10.1016/S1369-5274(03)00056-0.
79. Houfani, A. A., Anders, N., Spiess, A. C., Baldrian, P., & Benallaoua, S., (2020).
Insights from enzymatic degradation of cellulose and hemicellulose to fermentable
sugars: A review. Biomass Bioenergy, 134, 105481. https://doi.org/10.1016/j.biombioe.
2020.105481.
80. Anderson, C. T., (2016). We be jammin’: An update on pectin biosynthesis, trafficking
and dynamics. J. Exp. Bot., 67(2), 495–502. https://doi.org/10.1093/jxb/erv501.
81. Xiao, C., & Anderson, C. T., (2013). Roles of pectin in biomass yield and processing for
Author Copy
biofuels. Front. Plant Sci., 4. https://doi.org/10.3389/fpls.2013.00067.
82. Gummadi, S. N., & Panda, T., (2003). Purification and biochemical properties of
microbial pectinases: A review. Process Biochem., 38(7), 987–996. https://doi.org/
10.1016/S0032-9592(02)00203-0.
83. Pedrolli, D. B., Monteiro, A. C., Gomes, E., & Carmona, E. C., (2009). Pectin and pectinases:
Production, characterization and industrial application of microbial pectinolytic enzymes.
Open Biotechnol. J., 3(1), 9–18. https://doi.org/10.2174/1874070700903010009.
84. Singh, R. S., Singh, T., & Pandey, A., (2019). Microbial enzymes—An overview. In
Advances in Enzyme Technology (pp. 1–40). Elsevier. https://doi.org/10.1016/B978-0-
444-64114-4.00001-7.
85. Kashyap, D. R., Vohra, P. K., Chopra, S., & Tewari, R., (2001). Applications of
pectinases in the commercial sector: A review. Bioresour. Technol., 77(3), 215–227.
https://doi.org/10.1016/S0960-8524(00)00118-8.
86. Zhou, C. H., Xia, X., Lin, C. X., Tong, D. S., & Beltramini, J., (2011). Catalytic
conversion of lignocellulosic biomass to fine chemicals and fuels. Chem. Soc. Rev.,
40(11), 5588–5617. https://doi.org/10.1039/C1CS15124J.
87. Verbančič, J., Lunn, J. E., Stitt, M., & Persson, S., (2018). Carbon supply and the
regulation of cell wall synthesis. Mol. Plant, 11(1), 75–94. https://doi.org/10.1016/j.
molp.2017.10.004.
88. O’Leary, B. M., (2020). Another brick in the plant cell wall: Characterization of
Arabidopsis CSLD3 function in cell wall synthesis. Plant Cell, 32(5), 1359–1360.
https://doi.org/10.1105/tpc.20.00190.
89. Biswal, A. K., Atmodjo, M. A., Li, M., Baxter, H. L., Yoo, C. G., Pu, Y., Lee, Y. C., et al.,
(2018). Sugar release and growth of biofuel crops are improved by downregulation of
pectin biosynthesis. Nat. Biotechnol., 36(3), 249–257. https://doi.org/10.1038/nbt.4067.
90. Bhandari, S., Fujino, T., Thammanagowda, S., Zhang, D., Xu, F., & Joshi, C. P., (2006).
Xylem-specific and tension stress-responsive co-expression of Korrigan endoglucanase
and three secondary wall-associated cellulose synthase genes in aspen trees. Planta,
224(4), 828–837. https://doi.org/10.1007/s00425-006-0269-1.
91. Szyjanowicz, P. M. J., McKinnon, I., Taylor, N. G., Gardiner, J., Jarvis, M. C., & Turner,
S. R., (2004). The irregular xylem 2 mutant is an allele of Korrigan that affects the
secondary cell wall of Arabidopsis thaliana. Plant J., 37(5), 730–740.
92. McFarlane, H. E., Döring, A., & Persson, S., (2014). The cell biology of cellulose
synthesis. Annu. Rev. Plant Biol., 65(1), 69–94. https://doi.org/10.1146/annurev-arplant-
|050213-040240.
93. Van, M. A. J. A., Abbott, D. A., Bellissimi, E., Van, D. B. J., Kuyper, M., Luttik, M. A.
H., Wisselink, H. W., et al., (2006). Alcoholic fermentation of carbon sources in biomass
hydrolysates by Saccharomyces cerevisiae: Current status. Antonie Van Leeuwenhoek,
90(4), 391–418. https://doi.org/10.1007/s10482-006-9085-7.
Apple Academic Press
94. Schuster, B. G., & Chinn, M. S., (2013). Consolidated bioprocessing of lignocellulosic
feedstocks for ethanol fuel production. BioEnergy Res., 6(2), 416–435. https://doi.
org/10.1007/s12155-012-9278-z.
95. Plácido, J., & Capareda, S., (2015). Ligninolytic enzymes: A biotechnological alternative
for bioethanol production. Bioresour. Bioprocess, 2(1), 23. https://doi.org/10.1186/
s40643-015-0049-5.
96. Lin, Y., & Tanaka, S., (2006). Ethanol fermentation from biomass resources: Current state
and prospects. Appl. Microbiol. Biotechnol., 69(6), 627–642. https://doi.org/10.1007/
s00253-005-0229-x.
Author Copy
97. Amore, A., & Faraco, V., (2012). Potential of fungi as category I consolidated
bioprocessing organisms for cellulosic ethanol production. Renew. Sustain. Energy Rev.,
16(5), 3286–3301. https://doi.org/10.1016/j.rser.2012.02.050.
98. Ishola, M., Jahandideh, A., Haidarian, B., Brandberg, T., & Taherzadeh, M., (2013).
Simultaneous saccharification, filtration and fermentation (SSFF): A novel method for
bioethanol production from lignocellulosic biomass. Bioresour. Technol., 133, 68–73.
https://doi.org/10.1016/j.biortech.2013.01.130.
99. Salehi, J. G., & Taherzadeh, M. J., (2015). Advances in consolidated bioprocessing
systems for bioethanol and butanol production from biomass: A comprehensive review.
Biofuel Res. J., 2(1), 152–195. https://doi.org/10.18331/BRJ2015.2.1.4.
100. Zhang, J., & Lynd, L. R., (2010). Ethanol production from paper sludge by simultaneous
saccharification and co-fermentation using recombinant xylose-fermenting microor-
ganisms. Biotechnol. Bioeng., 107(2), 235–244. https://doi.org/10.1002/bit.22811.
101. Sivasubramanian, V., (2018). Bioprocess Engineering for a Green Environment. CRC
Press.
102. Olson, D. G., McBride, J. E., Joe, S. A., & Lynd, L. R., (2012). Recent progress in consoli-
dated bioprocessing. Curr. Opin. Biotechnol., 23(3), 396–405. https://doi.org/10.1016/j.
copbio.2011.11.026.
103. Kumagai, A., Kawamura, S., Lee, S. H., Endo, T., Rodriguez, M., & Mielenz, J. R.,
(2014). Simultaneous saccharification and fermentation and a consolidated bioprocessing
for hinoki cypress and eucalyptus after fibrillation by steam and subsequent wet-disk
milling. Bioresour. Technol., 162, 89–95. https://doi.org/10.1016/j.biortech.2014.03.110.
104. Kamei, I., Hirota, Y., Mori, T., Hirai, H., Meguro, S., & Kondo, R., (2012). Direct
ethanol production from cellulosic materials by the hypersaline-tolerant white-rot
fungus Phlebia Sp. MG-60. Bioresour. Technol., 112, 137–142. https://doi.org/10.1016/j.
biortech.2012.02.109.
105. Hasunuma, T., & Kondo, A., (2012). Consolidated bioprocessing and simultaneous
saccharification and fermentation of lignocellulose to ethanol with thermotolerant
yeast strains. Process Biochem., 47(9), 1287–1294. https://doi.org/10.1016/j.procbio.
2012.05.004.
106. Den, H. R., Van, R. E., Rose, S. H., Görgens, J. F., & Van, Z. W. H., (2015). Progress
and challenges in the engineering of non-cellulolytic microorganisms for consolidated
bioprocessing. Curr. Opin. Biotechnol., 33, 32–38. https://doi.org/10.1016/j.copbio.2014.
10.003.
107. Sarkar, N., Ghosh, S. K., Bannerjee, S., & Aikat, K., (2012). Bioethanol production from
agricultural wastes: An overview. Renew. Energy, 37(1), 19–27. https://doi.org/10.1016/
j.renene.2011.06.045.
108. Liers, C., Arnstadt, T., Ullrich, R., & Hofrichter, M., (2011). Patterns of lignin degradation
Apple Academic Press
Author Copy
Biophys., 574, 66–74. https://doi.org/10.1016/j.abb.2015.01.018.
111. Fernández-Fueyo, E., Castanera, R., Ruiz-Dueñas, F. J., López-Lucendo, M. F., Ramírez,
L., Pisabarro, A. G., & Martínez, A. T., (2014). Ligninolytic peroxidase gene expression
by Pleurotus ostreatus: Differential regulation in lignocellulose medium and effect of
temperature and PH. Fungal Genet. Biol., 72, 150–161. https://doi.org/10.1016/j.fgb.
2014.02.003.
112. Kumar, A., & Chandra, R., (2020). Ligninolytic enzymes and its mechanisms for
degradation of lignocellulosic waste in the environment. Heliyon, 6(2), e03170. https://
doi.org/10.1016/j.heliyon.2020.e03170.
113. Martínez, Á. T., Speranza, M., Ruiz-Dueñas, F. J., Ferreira, P., Camarero, S., Guillén, F.,
Martínez, M. J., et al., (2005). Biodegradation of lignocellulosics: Microbial, chemical,
and enzymatic aspects of the fungal attack of lignin. Int. Microbiol., 8(3), 195–204.
https://doi.org/10.2436/im.v8i3.9526.
114. Martínez, Á. T., Ruiz-Dueñas, F. J., Martínez, M. J., Del, R. J. C., & Gutiérrez, A.,
(2009). Enzymatic delignification of plant cell wall: From nature to mill. Curr. Opin.
Biotechnol., 20(3), 348–357. https://doi.org/10.1016/j.copbio.2009.05.002.
Author Copy
LEOPOLDO JAVIER RÍOS GONZÁLEZ,2
MIGUEL ANGEL MEDINA MORALES,2
MÓNICA MARGARITA RODRÍGUEZ GARZA,2
ILEANA MAYELA MARÍA MORENO DÁVILA,2
THELMA KARINA MORALES MARTÍNEZ,3 and
JOSÉ ANTONIO RODRÍGUEZ DE LA GARZA2
1
Botany Department, Agronomic Division, Antonio Narro Autonomous
Agrarian University, Mexico
2
Department of Biotechnology, School of Chemistry, Autonomous
University of Coahuila, Mexico, Phone: +52 (844) 415-5752–Ext 120,
E-mail: antonio.rodriguez@uadec.edu.mx (J. A. R. D. L. Garza)
3
Bioprocess and Microbial Biochemistry Group, School of Chemistry,
Autonomous University of Coahuila, Mexico
ABSTRACT
Microbial electrosynthesis has gained much attention in the last decade in the
scientific community due to that this novel process allows the production of
commodity chemicals sustainably. MES (Microbial electrosynthesis cell), as
commonly referred to as this technology, is a bioelectrochemical system that
utilizes the reducing power generated from the anodic oxidation to produce
high value-added products in the cathode. When coupled with a renewable
source of electricity and organic or inorganic substances to produce biofuels,
such as alcohols (ethanol, butanol, etc.), significant advantages stand out,
such as the avoidance of fossil fuels to provide the energy that drives the
process and food crops that serve as feedstock. The present chapter will
give an insight into the novel perspectives of bioethanol production through
bioelectrosynthesis as well as the more recent and relevant developments in
MES technology for this purpose.
Apple Academic Press
16.1 INTRODUCTION
2020 has been a year that has not only have cause several health issues all
around the world but also has shown the fragility of the global economy
[1] and its dependence on fossil fuel. This dependence has been more of
Author Copy
a reminder of the need to search, and develop sustainable fuel production
processes to fulfill our energy needs and, additionally, options that have
higher market stability than oil prices [2–5]. In this regard, biofuels were
considered an alternative to gradually replace fossil fuel over the next 30
years; however, many concerns have arisen mainly due to the feedstocks
used to produce biofuels, generating food security issues all around the
globe. In the past decade, it has been clear that future energy demand will be
met by a combination of renewable energy (solar, hydro, thermal, biofuels,
and eolic) and fossil fuels [6–8].
As stated initially, this year (2020) has presented several challenges;
however, also it has opened new opportunities. The Contraction in global
oil consumption has opened the windows for the renewable energy market;
additionally, the energy market is being subjected to pressure to lower its gas
emissions to fulfill the objectives of reverting or at least tackling the climate
change issue [2–5].
The use of renewable energy, combined with bioenergy, will allow us
to be more independent in energy matters and fight climate change. Among
bioenergy, bio-alcohols as high value-added products from agro-industrial
waste as an organic carbon source are well established in the energy market.
Among bioalcohols, bioethanol is by far, the most produced biofuel in the
globe, being Brazil and the United States the dominant players in bioethanol
production, representing almost 80% of the global output. However, as
mentioned earlier, conflicts over food supply and land use have made its
production and utilization a controversial issue. To overcome this issue, in
the last decade, second-generation bioethanol production technology has
focused on the use of non-edible (human consumption) biomass sources,
using a wide range of feedstocks, such as LCB, energy crops, cellulosic
residues, and, particularly, wastes [9–11]. The quality of bioethanol produced
is majorly dependent on the production routes. Bioethanol production, in
Author Copy
issues mentioned before. Although ethanol can be produced directly from
CO2, its production from acetate is thermodynamically and energetically
more favorable, providing that the undissociated acid exists in a slightly
acidic medium [13–15]. Still, ethanol production through MES is far from
becoming a feasible process, yet as titers and efficiencies are relatively
low. The following book chapter will be focused on describing briefly on
describing the current development on BES and its multiple application,
followed by exploring the use of BES for ethanol production by means of
MES. It will be discussed how can BES compete, in terms of energy usage,
with conventional bioethanol production technologies by exploiting the
energy content of some of the chemicals present in the waste biomass and
CO2 waste streams.
for CO2 bioconversion to methane was discovered as well, this is, “elec-
tromethanogenesis” that we now know [15, 31–37]. Based on these initial
efforts, a great deal of MEC configurations with different purposes have been
designed and proposed very recently, such as microbial electrodialysis cell
(MEDC), microbial reverse electrodialysis electrolysis cell (MREC), micro-
bial desalination cell (MDC), etc., substantially extending the scope of MEC
applications. From then onwards, a new era belonging to MEC is approaching
Author Copy
(Figure 16.2) [29, 30].
FIGURE 16.1 Number of published journal articles on MES containing the phrases
“microbial fuel cell;” “microbial electrolysis cell;” “microbial electrosynthesis;” or “microbial
desalination cell.”
Source: Adapted from: 2013–2019 PubMed Central.
Source: Based on figure presented by Wang and Ren [22].
The different types of BES can distinguish from one another by the main func-
tion and the output of the BES. For example, a microbial fuel cell (MFC) is the
most characteristic type of BES, whose main function is the generation of direct
electricity. When an external power supply is added to an MFC reactor to reduce
the potential of the cathode, the system becomes a microbial electrolysis cell
(MEC), where hydrogen gas and other products are generated [22]. Addition-
ally, the potential input required in a MEC can be provided by a MFC so that
hydrogen can be produced directly from biomass and waste sources [38]. If the
Apple Academic Press
Author Copy
tion cell or MDS [22]. Nutrient removal or recovery can also be carried out by
a BES, in the case of nitrogen, it can be removed through bioelectrochemical
denitrification or recovered vía ammonium migration driven by electricity
generation (Figure 16.3) [20, 39, 40].
FIGURE 16.2 Overview of various types of bioelectrochemical systems and main products.
volatile fatty acids, biomass, and carbon dioxide (CO2) [45]. This type of
BES has the potential of a dual application, treating different types of waste
streams, such as wastewater, and during this process, an electric current will
be produced. In this approach, it generates an alternative energy source from
Apple Academic Press
Author Copy
Author Copy
the additional applied voltage that helps overcome the cathode limitations.
The energy required for MEC operation can also be provided by another
separate MFC as a power source [28, 57–60]. In a MEC with bioanode and
biocathode, expensive metals like platinum are not required as catalyst, and
preferably, the enrichment of microbes on the carbon cathode decreases the
start-up time and produces comparable current densities to those of bioanode
[61–67]. Furthermore, the hydrogen synthesized in MEC can also drive the
biochemical production of other high value-added chemicals [22, 63, 68, 67].
Author Copy
carbon and electron flow is channeled to the production of extracellular end
products rather than to the microbial growth and biomass production [15, 33,
49]. The thermodynamics of ethanol production by CO2 or acetate reduction
indicates that lower pH and excess H2 can shift the production to ethanol.
The selectivity of the products of microbial CO2 reduction has been shown to
be limited by micro-elements of trace metal and vitamin components due to
their effect on the activity of NADPH/NADP and Acetyl-CoA enzymes [15,
32, 78, 79]. By limiting micronutrient and trace elements (namely pantothenic
acid and cobalt), higher concentrations of ethanol can be produced rather
than acetate, suggesting that larger multicarbon compounds can be produced
by means of applying strategies like those use in the gas fermentation process
in where gaseous feedstocks such as CO, CO2, syngas, CH4 or biogas, into
platform chemicals, fuels, polymers, etc. Thermodynamically, the reduction
of acetate to ethanol is more favorable than the reduction of bicarbonate
to ethanol and this has been reported when acetate accumulated and goes
through further reduction (Figure 16.4). However, ethanol concentrations
achieved are quite low [14, 34, 49, 56, 73, 80, 81].
One aspect that still needs further study is the mechanism of bacterial
interaction with solid electrodes for CO2 reduction. In this regard, the
overpotential of the electrode will create conditions that much likely require
a lower potential in the cathode. The necessary overpotential so that CO2
reduction can take place in the cathode will require much lower cathode
potential than the hydrogen evolution reaction (HER) [82]. Additionally,
in the case of using a mixed culture, most likely, the HER will take place
first, followed by CO2 reduction mediated by H2 [6, 34, 67, 69, 72]. On the
other hand, when direct electron transfer takes place, it could be possible to
achieve higher production rates of the targeted compound. Cytochrome-c
for extracellular electron transfer on homoacetogenic bacteria has shown the
potential of different of using multiple mechanisms of electron transfer, and
this will depend of the biocatalyst chosen (Figure 16.5). In most of the litera-
ture, the bioelectrochemical reduction of CO2 has been reported to take place
at lower cathode potentials than the theoretical HER potential (0.61 V when
reference electrode is Ag/AgCl). MES experiments showed a considerable
Apple Academic Press
production of acetate only when the potentials were more negative than 0.8
V (using as reference electrode Ag/AgCl) and whenever the potentials were
above or equal to 0.8 V, acetate, and butyrate were degraded instead of CO2
reduction. In this regard, acetate production occurs only below 0.9 V, in
this case H2 should mediate the CO2 reduction reaction. Bioelectrocatalsys
combined with the HER can take place when higher current densities (up to
10 ampers per m2) are obtained at a potential of 0.9 to 1 V and when abiotic
Author Copy
cathode is being used. However, when the biocathode is a mixed culture, it
would not be clear what kind of mechanisms for electron transfer will be
taking place [32, 33, 49, 52, 62, 76, 78, 81, 83–90].
FIGURE 16.4 Schematic of the Wood-Ljungdahl pathway for CO2 reduction for ethanol
production.
16.8 CONCLUSIONS
MES can offer a wide range of opportunities that go from, energy production
from wastes, to the production of high added value products from inorganic
sources. MES technology offers new approaches that other heavily
Author Copy
FIGURE 16.5 Extracellular electron transfer mechanism of Geobacter sulfurreducens [41].
solutions, etc.). Inexpensive and abundant carbon sources including CO2 and
waste streams can be reutilized for bioproduction with the input of electrical
energy only (renewable energy). In addition, MES offers unique possibilities
to promote cell growth and the controlling of redox state in bioprocesses by
Apple Academic Press
Author Copy
KEYWORDS
REFERENCES
1. Atalan, A., (2020). Is the lockdown important to prevent the COVID-9 pandemic?
Effects on psychology, environment and economy-perspective. Annals of Medicine and
Surgery, 56, 38–42. Elsevier. doi: 10.1016/j.amsu.2020.06.010.
2. Prest, B. C., (2018). Explanations for the 2014 oil price decline: Supply or demand?.
Energy Economics, 74, 63–75. doi: 10.1016/j.eneco.2018.05.029.
3. Singh, V. K., Nishant, S., & Kumar, P., (2018). Dynamic and directional network
connectedness of crude oil and currencies: Evidence from implied volatility. Energy
Economics, 76, 48–63. Elsevier B.V. doi: 10.1016/j.eneco.2018.09.018.
4. Wu, X. F., & Chen, G. Q., (2019). Global overview of crude oil use: From source to sink
through inter-regional trade. Energy Policy, 128, 476–486. doi: 10.1016/j.enpol. 2019.01.022.
5. Go, Y. H., & Lau, W. Y., (2020). The impact of global financial crisis on informational
efficiency: Evidence from price-volume relation in crude palm oil futures market. Journal
of Commodity Markets, 17, 100081. Elsevier Ltd. doi: 10.1016/j.jcomm.2018.10.003.
6. Braden, D. J., (2005). Fuel Cell Grade Hydrogen Production from Steam Reforming of
Bio-Ethanol Over Co-based Catalyst: An Investigation of Reaction Networks and Active
Sites. Metallum. Com. Br, The Ohio S.
7. Jadhav, D. A., Ghosh, R. S., & Ghangrekar, M. M., (2017). Third generation in
bio-electrochemical system research - a systematic review on mechanisms for recovery
of valuable by-products from wastewater. Renewable and Sustainable Energy Reviews,
76, 1022–1031. doi: 10.1016/j.rser.2017.03.096.
Apple Academic Press
Author Copy
11. Zhao, J., et al., (2020). Bioconversion of industrial hemp biomass for bioethanol
production: A review. Fuel, 281. doi: 10.1016/j.fuel.2020.118725.
12. Aditiya, H. B., et al., (2016). Second-generation bioethanol production: A critical review.
Renewable and Sustainable Energy Reviews, 66, 631–653. Elsevier. doi: 10.1016/j.rser.
2016.07.015.
13. Warriner, K., et al., (2002). Modified microelectrode interfaces for in-line electrochemical
monitoring of ethanol in fermentation processes. Sensors and Actuators, B: Chemical,
84(2, 3), 200–207. doi: 10.1016/S0925-4005(02)00025-4.
14. Pagnoncelli, K. C., et al., (2018). Ethanol generation, oxidation and energy production
in a cooperative bioelectrochemical system. Bioelectrochemistry, 122, 11–25. Elsevier
B.V. doi: 10.1016/j.bioelechem.2018.02.007.
15. Prévoteau, A., et al., (2020). Microbial electrosynthesis from CO2: Forever a promise?.
Current Opinion in Biotechnology, 62, 48–57. doi: 10.1016/j.copbio.2019.08.014.
16. Potter, M. C., (1910). On the Difference of Potential Due to the Vital Activity of
Microorganisms. Proc Univ Durham Phil Soc.
17. Morris, J. M., et al., (2009). Microbial Fuel Cell in Enhancing Anaerobic Biodegradation
of Diesel. 146, 161–167. doi: 10.1016/j.cej.2008.05.028.
18. Morris, J. M., & Jin, S., (2012). Enhanced biodegradation of hydrocarbon-contaminated
sediments using microbial fuel cells. Journal of Hazardous Materials, 213, 214,
474–477. Elsevier B.V. doi: 10.1016/j.jhazmat.2012.02.029.
19. Kronenberg, M., et al., (2017). Biodegradation of polycyclic aromatic hydrocarbons:
Using microbial bioelectrochemical systems to overcome an impasse. Environmental
Pollution, 231, 509–523. Elsevier Ltd. doi: 10.1016/j.envpol.2017.08.048.
20. Nancharaiah, Y. V., Mohan, S. V., & Lens, P. N. L., (2019). Removal and recovery
of metals and nutrients from wastewater using bioelectrochemical systems. Microbial
Electrochemical Technology. Elsevier B.V. doi: 10.1016/b978-0-444-64052-9.00028-5.
21. Cohen, B., (1931). The bacterial culture as an electrical half-cell. Journal of Bacteriology,
21, 18–19.
22. Wang, H., & Ren, Z. J., (2013). A comprehensive review of microbial electrochemical
systems as a platform technology. Biotechnol. Adv., 31, 1796–1807.
23. Liu, H., Grot, S., & Logan, B. E., (2005). Electrochemically assisted microbial production
of hydrogen from acetate. Environmental Science and Technology, 39(11), 4317–4320.
doi: 10.1021/es050244p.
24. Kadier, A., et al., (2016). A comprehensive review of microbial electrolysis cells (MEC)
reactor designs and configurations for sustainable hydrogen gas production. Alexandria
Author Copy
Microbial Electrolysis Cells: Studies Towards its Scaling-up, 192. Available at: https://
www.tdx.cat/bitstream/handle/10803/283897/nmip.pdf;jsessionid=EB7C604404BCA
BEE1D791B0298D3CA2D?sequence=1 (accessed on 08 December 2021).
29. Zhang, Y., et al., (2019). Microbial fuel cell hybrid systems for wastewater treatment and
bioenergy production: Synergistic effects, mechanisms and challenges. Renewable and
Sustainable Energy Reviews, 103, 13–29. Elsevier Ltd. doi: 10.1016/j.rser.2018.12.027.
30. Chu, N., et al., (2020). Microbial electrochemical platform for the production of
renewable fuels and chemicals. Biosensors and Bioelectronics, 150, 111922. Elsevier
B.V. doi: 10.1016/j.bios.2019.111922.
31. Thygesen, A., et al., (2010). Integration of microbial electrolysis cells (MECs) in the
biorefinery for production of ethanol, H2 and phenolics. Waste and Biomass Valorization,
1(1), 9–20. doi: 10.1007/s12649-009-9007-9.
32. Arends, J. B. A., et al., (2017). Continuous long-term electricity-driven bioproduction of
carboxylates and isopropanol from CO2 with a mixed microbial community. Journal of
CO2 Utilization, 20, 141–149. Elsevier. doi: 10.1016/j.jcou.2017.04.014.
33. Bajracharya, S., Van, D. B. B., et al., (2017). In situ acetate separation in microbial
electrosynthesis from CO2 using ion-exchange resin. Electrochimica Acta, 237, 267–275.
Elsevier Ltd. doi: 10.1016/j.electacta.2017.03.209.
34. Anwer, A. H., et al., (2019). Development of novel MnO2 coated carbon felt cathode for
microbial electroreduction of CO2 to biofuels. Journal of Environmental Management,
249, 109376. Elsevier. doi: 10.1016/j.jenvman.2019.109376.
35. Bian, B., et al., (2020). Microbial electrosynthesis from CO2: Challenges, opportunities
and perspectives in the context of circular bioeconomy. Bioresource Technology, 302,
122863. Elsevier Ltd. doi: 10.1016/j.biortech.2020.122863.
36. Singh, S., Noori, M. T., & Verma, N., (2020). Efficient bio-electroreduction of CO2
to formate on a iron phthalocyanine-dispersed CDC in microbial electrolysis system.
Electrochimica Acta, 338, 135887. Elsevier Ltd. doi: 10.1016/j.electacta.2020.135887.
37. Xiao, S., et al., (2020). Hybrid microbial photoelectrochemical system reduces CO2 to
CH4 with 1.28% solar energy conversion efficiency. Chemical Engineering Journal,
390, 124530. Elsevier B.V. doi: 10.1016/j.cej.2020.124530.
38. Pant, D., Singh, A., Van-Bogaert, G., Alvarez-Gallego, Y., Diels, L., & Vanbroekhoven,
K., (2010). An introduction to the life cycle assessment (LCA) of bioelectrochemical
systems (BES) for sustainable energy and product generation: Relevance and key
aspects. Energy Rev., 1305–13.
39. Zhang, F., Li, J., & He, Z., (2014). A new method for nutrient removal and recovery
from wastewater using a bioelectrochemical system. Bioresour. Technol., 166, 630–634.
40. Zhang, G., Zhou, Y., & Yang, F., (2019). Hydrogen production from microbial fuel
cells-ammonia electrolysis cell coupled system fed with landfill leachate using Mo2
Apple Academic Press
Author Copy
in Microbial Fuel Cells, 1, 7–15.
44. Santoro, C., et al., (2017). Microbial fuel cells: From fundamentals to applications. A
review. Journal of Power Sources, 356, 225–244. Elsevier B.V. doi: 10.1016/j.jpowsour.
2017.03.109.
45. Velasquez-Orta, S. B., Head, I. M., Curtis, T. P., & Scott, K., (2011). Factors affecting
current production in microbial fuel cells using different industrial wastewaters. Biore-
source Technol., 5105–5012.
46. Zhou, M., He, H., & Jin, T., & Wang, H., (2012). Power generation enhancement in
novel microbial carbon capture cells with immobilized Chlorella vulgaris. J. Power
Sources, 216–19.
47. Venkata-Mohan, S., Velvizhi, G., Vamshi-Krishna, K., & Lenin-Babu, M. (2014).
Microbial catalyzed electrochemical systems: A biofactory with multi-facet applications.
Bioresource Technol., 355–364.
48. Eerten-Jansen, M. C. A. A. V, et al., (2013). Microbial community analysis of a
methane-producing biocathode in a bioelectrochemical system. Archaea, 2013. doi:
10.1155/2013/481784.
49. Gildemyn, S., et al., (2015). Integrated production, extraction, and concentration of
acetic acid from CO2 through microbial electrosynthesis. Environmental Science and
Technology Letters, 2(11), 325–328. doi: 10.1021/acs.estlett.5b00212.
50. Sadhukhan, J., et al., (2016). A critical review of integration analysis of microbial
electrosynthesis (MES) systems with waste biorefineries for the production of biofuel
and chemical from reuse of CO2. Renewable and Sustainable Energy Reviews, 56,
116–132. Elsevier. doi: 10.1016/j.rser.2015.11.015.
51. Bajracharya, S., Srikanth, S., et al., (2017a). Biotransformation of carbon dioxide in
bioelectrochemical systems: State of the art and future prospects. Journal of Power
Sources, 356, 256–273. Elsevier B.V. doi: 10.1016/j.jpowsour.2017.04.024.
52. Bajracharya, S., Vanbroekhoven, K., et al., (2017). Bioelectrochemical conversion of
CO2 to chemicals: CO2 as a next-generation feedstock for electricity-driven bioproduc-
tion in batch and continuous modes. Faraday Discussions, 202, 433–449. doi: 10.1039/
c7fd00050b.
53. Nelabhotla, A. B. T., Bakke, R., & Dinamarca, C., (2019). Performance analysis
of biocathode in bioelectrochemical CO2 reduction. Catalysts, 9(8). doi: 10.3390/
catal9080683.
for sequential reduction of CO2 to methane. Fuel, 220, 8–13. Elsevier. doi: 10.1016/j.
fuel.2018.01.141.
56. Kokkoli, A., Zhang, Y., & Angelidaki, I., (2018). Microbial electrochemical separation
of CO2 for biogas upgrading. Bioresource Technology, 247, 380–386. Elsevier Ltd. doi:
10.1016/j.biortech.2017.09.097.
57. Montpart, N., et al., (2016). Low-cost fuel-cell based sensor of hydrogen production
in lab-scale microbial electrolysis cells. International Journal of Hydrogen Energy,
41(45), 20465–20472. doi: 10.1016/j.ijhydene.2016.09.169.
58. Fischer, F., (2018). Photoelectrode, photovoltaic and photosynthetic microbial fuel cells.
Author Copy
Renewable and Sustainable Energy Reviews, 90, 16–27. Elsevier Ltd. doi: 10.1016/j.
rser.2018.03.053.
59. Huarachi-Olivera, R., et al., (2018). Bioelectrogenesis with microbial fuel cells (MFCs)
using the microalga Chlorella vulgaris and bacterial communities. Electronic Journal
of Biotechnology, 31, 34–43. Elsevier España, S.L.U. doi: 10.1016/j.ejbt.2017.10.013.
60. Zhang, J., Yuan, H., Abu-Reesh, I. M., He, Z., & Yuan, C., (2019). Life cycle environmental
impact comparison of bioelectrochemical systems for wastewater treatment. Procedia
CIRP, 80, 382–388.
61. Butti, S. K., et al., (2016). Microbial electrochemical technologies with the perspective
of harnessing bioenergy: Maneuvering towards upscaling. Renewable and Sustainable
Energy Reviews, 53, 462–476. Elsevier. doi: 10.1016/j.rser.2015.08.058.
62. Bajracharya, S., Yuliasni, R., et al., (2017). Long-term operation of microbial
electrosynthesis cell reducing CO2 to multi-carbon chemicals with a mixed culture
avoiding methanogenesis. Bioelectrochemistry, 113, 26–34. Elsevier B.V. doi: 10.1016/j.
bioelechem.2016.09.001.
63. Kumar, G., et al., (2017). A review on bio-electrochemical systems (BESs) for the
syngas and value-added biochemicals production. Chemosphere, 177, 84–92. Elsevier
Ltd. doi: 10.1016/j.chemosphere.2017.02.135.
64. Li, J., et al., (2017). Uneven biofilm and current distribution in three-dimensional
macroporous anodes of bio-electrochemical systems composed of graphite electrode
arrays. Bioresource Technology, 228, 25–30. Elsevier Ltd. doi: 10.1016/j.biortech.
2016.12.092.
65. Zou, S., & He, Z., (2018). Efficiently “pumping out” value-added resources from
wastewater by bioelectrochemical systems: A review from energy perspectives. Water
Research, 131, 62–73. Elsevier Ltd. doi: 10.1016/j.watres.2017.12.026.
66. Yang, E., et al., (2019). Critical review of bioelectrochemical systems integrated with
membrane-based technologies for desalination, energy self-sufficiency, and high-
efficiency water and wastewater treatment. Desalination, 452, 40–67. Elsevier. doi:
10.1016/j.desal.2018.11.007.
67. Luo, S., & He, Z., (2020). Resource Recovery from Wastewater by Bioelectrochemical
Systems, 183–200. Elsevier Inc. doi: 10.1016/B978-0-12-816204-0.00009-6.
68. Bajracharya, S., et al., (2016). An overview on emerging bioelectrochemical systems
(BESs): Technology for sustainable electricity, waste remediation, resource recovery,
chemical production and beyond. Renewable Energy, 98, 153–170. Elsevier Ltd. doi:
10.1016/j.renene.2016.03.002.
69. Ramírez-Vargas, C. A., et al., (2018). Microbial electrochemical technologies
for wastewater treatment: Principles and evolution from microbial fuel cells to
Apple Academic Press
Author Copy
watres.2018.10.092.
73. Schievano, A., Pant, D., & Puig, S., (2019). Editorial: Microbial synthesis,
gas-fermentation and bioelectroconversion of CO2 and other gaseous streams. Frontiers
in Energy Research, 7, 1–4. doi: 10.3389/fenrg.2019.00110.
74. Jin, S., & Fallgren, P. H., (2014). Feasibility of using bioelectrochemical systems for
bioremediation. Microbial Biodegradation and Bioremediation, 389–405. Elsevier. doi:
10.1016/B978-0-12-800021-2.00016-9.
75. Li, X., Angelidaki, I., & Zhang, Y., (2018). Salinity-gradient energy-driven microbial
electrosynthesis of value-added chemicals from CO2 reduction. Water Research, 142,
396–404. Elsevier Ltd. doi: 10.1016/j.watres.2018.06.013.
76. Irfan, M., et al., (2019). Direct microbial transformation of carbon dioxide to value-
added chemicals: A comprehensive analysis and application potentials. Bioresource
Technology, 288, 121401. Elsevier. doi: 10.1016/j.biortech.2019.121401.
77. Srivastava, R. K., (2019). Bio-energy production by contribution of effective and
suitable microbial system. Materials Science for Energy Technologies, 2(2), 308–318.
KeAi Communications Co., Ltd. doi: 10.1016/j.mset.2018.12.007.
78. Ragsdale, S. W., & Pierce, E., (2008). Acetogenesis and the wood-ljungdahl pathway
of CO2 fixation. Biochimica et Biophysica Acta - Proteins and Proteomics, 1784(12),
1873–1898. doi: 10.1016/j.bbapap.2008.08.012.
79. Awate, B., et al., (2017). Stimulation of electro-fermentation in single-chamber
microbial electrolysis cells driven by genetically engineered anode biofilms. Journal
of Power Sources, 356, 510–518. Elsevier B.V. doi: 10.1016/j.jpowsour.2017.02.053.
80. Speers, A. M., Young, J. M., & Reguera, G., (2014). Fermentation of glycerol into ethanol
in a microbial electrolysis cell driven by a customized consortium. Environmental
Science and Technology, 48(11), 6350–6358. doi: 10.1021/es500690a.
81. Gavilanes, J., Reddy, C. N., & Min, B., (2019). Microbial electrosynthesis of bioalcohols
through reduction of high concentrations of volatile fatty acids. Energy and Fuels, 33(5),
4264–4271. doi: 10.1021/acs.energy fuels.8b04215.
82. Doğan, H. Ö., (2019). Ethanol electro-oxidation in alkaline media on Pd/electrodeposited
reduced graphene oxide nanocomposite modified nickel foam electrode. Solid-State
Sciences, 98. doi: 10.1016/j.solidstatesciences.2019.106029.
83. Steinbusch, K. J. J., et al., (2010). Bioelectrochemical ethanol production through
mediated acetate reduction by mixed cultures. Environmental Science and Technology,
44(1), 513–517. doi: 10.1021/es902371e.
84. Andersen, S. J., et al., (2014). Electrolytic membrane extraction enables the production
of fine chemicals from biorefinery side streams. Environmental Science and Technology,
48(12), 7135–7142. doi: 10.1021/es500483w.
85. Ammam, F., et al., (2016). Effect of tungstate on acetate and ethanol production by
Apple Academic Press
Author Copy
88. Paiano, P., et al., (2019). Electro-fermentation and redox mediators enhance glucose
conversion into butyric acid with mixed microbial cultures. Bioelectrochemistry, 130,
107333. Elsevier B.V. doi: 10.1016/j.bioelechem.2019.107333.
89. Im, H. S., et al., (2019). Isolation of novel CO converting microorganism using zero-
valent iron for a bioelectrochemical system (BES). Biotechnology and Bioprocess
Engineering, 24(1), 232–239. doi: 10.1007/s12257-018-0373-7.
90. Haavisto, J. M., et al., (2020). The effect of start-up on energy recovery and compositional
changes in brewery wastewater in bioelectrochemical systems. Bioelectrochemistry,
132, 107402. Elsevier LTD. doi: 10.1016/j.bioelechem.2019.107402.
Author Copy
M. J. CASTRO-ALONSO,1 R. T. HILARES,2 P. R. F. MARCELINO,1
C. A. PRADO,1 M. M. CAMPOS,1 A. S. CARDOSO,1 J. C. SANTOS,3
and S. S. DA SILVA1
1
Bioprocesses and Sustainable Products Laboratory,
Department of Biotechnology, Engineering School of Lorena,
University of São Paulo (EEL-USP), Lorena–12.602.810, SP, Brazil,
E-mails: salvador.sanchez@usp.br;
sanchezmunoz.ssm@gmail.com (S. Sánchez-Muñoz)
2
Material Laboratory, Catholic University of Santa Maria (UCSM),
Yanahuara, AR–04013, Peru
Bioprocesses, Biopolymers, Simulation, and Modeling Laboratory,
3
ABSTRACT
Environmental and economic impacts caused by the use of fossil fuels have
been demanding the development of renewable energy technologies based on
low carbon fuels. In the last times, large-scale investments in the production of
biofuels have increased in many countries intending to achieve the Sustainable
Development Scenario. However, policies, economic competitors, regional
feedstocks availability, technological challenges of biomass conversion, and
bottlenecks for current and future biofuels productions, represent key issues
to be considered for their incorporation into the transport sector.
17.1 INTRODUCTION
In the transport sector, there is the dominance of the use of fossil fuels that are
derived from oil sources. With global economic development, the demand
Apple Academic Press
for fuels is growing proportionately. The increasing use of these fuels has
been pointed to as responsible for the occurrence of climate changes due to
an increased concentration of CO2, the main greenhouse gas (GHG), in the
atmosphere. Reports inform us that the transport sector consumes 30% of
global energy and is responsible for 21% of the annual GHG emissions [1].
As a consequence of environmental concerns, the search for sustainable
fuels (biofuels) production with low CO2 emission and minimal climate
Author Copy
changes has grown. In the world, the countries with the highest production
of biofuels are the USA, Brazil, and China, and it is expected that these
countries will double or triple their consumption of biofuels by 2030 [2, 3].
Worldwide, the biofuels most commonly used in motor vehicles,
marine vessels, and aviation are bioethanol, biodiesel, and biogas [4, 5].
Bioethanol and biodiesel are mainly produced by USA and Brazil, while
for biogas, the largest production is observed in Europe. Even though,
policies and technologies need to be improved to allow and encourage the
application of these biofuels in transportation. For this, government projects
and programs are developed to intensify the production and use of these
sustainable fuels. Also, financial subsidies are provided to manufacturers
aiming for the production of sustainable supplies from available biomass to
support bioeconomy development [6]. Due to these financial subsidies and
government programs, biofuels classified as the first-generation dominate
the sector of renewable fuels applied to transportation. Their production is at
an advanced stage, from the processing of the biomasses involved until the
industrial infrastructure. However, aspects related to land availability and
the competition of feedstocks with the food sector, make second and third-
generation technologies gaining attention worldwide [7, 8].
In this context, the perspectives and challenges faced for the production
and application of biofuels in the transport sector will be discussed in this
chapter.
Biofuels can replace conventional fuels derived from oil, such as gasoline
and diesel in the transportation sector [187]. The obtaining processes of
those renewable forms of fuel and their use must be environmentally friendly
[9–11]. Available biofuels are bioethanol, biodiesel, biogas (biomethane),
and biohydrogen [12, 13].
Biofuels are classified as first-generation (1G), second-generation (2G),
Apple Academic Press
and third-generation (3G) [14–19, 188]. This classification relates to the type
of feedstocks used in their production, for example, 1G biofuels have as
feedstock energy crops like rice, wheat, corn, sugarcane, and also vegetable
oils, among others [20]. On the other hand, 2G biofuels are obtained from
wood residues, inedible vegetable oil, non-food crops, and crop byproducts
such as wheat straw, corn straw, sugarcane bagasse [21]. In the 3G biofuels,
the biomass of organisms is used as raw material, such as microalgae and
Author Copy
seaweed [22, 23]. Currently, industrial global production is led by 1G biofuels
from traditional biomass, and it has been estimated that this technology will
still be responsible for the development of biofuels for years to come. For
example, bioethanol production is expected to use 14% and 24% of global
corn and sugarcane production by 2028 [24].
In 2018, about 14% of the energy in the world came from renewable
sources, and the other 86% corresponds to non-renewable energy, including
coal, crude oil, natural gas [25]. From whole renewable energy, approximately
3.7% was directed to the transport sector [26]. Thus, biofuels contribute only
with a small part of the energy supply in transportation, but their use has
been increasing in parallel with the growth of the vehicle fleet [4, 27].
In 2019, the world production of biofuels was approximately 154.5 billion
liters, and it is expected to increase by 25% during the 2019–2024 period,
reaching about 190 billion liters of biofuels [28]. USA and Brazil were the
largest producers, generating around 66.8 and 37 billion liters, respectively
[25, 29]. Among the biofuels available in top producer countries for trans-
portation, here will be highlighted biogas, biodiesel, and bioethanol [4].
Biogas corresponds to a mixture of combustible gases that can be produced
by AD of manure, farm wastes, wastewater, and other residues [30–32]. It
is used for the domestic generation of energy/heat, but also can be used as
a motor fuel in traditional vehicle designs [33, 34]. Europe, Asia, America,
and Oceania are the main biogas producers, as shown in Figure 17.1 [35].
This biofuel is mostly used in energy conversion into electricity, but its use
in vehicles for transportation still has limitations [32, 36]. Although, when
biogas is upgraded to biomethane, its use as a transport fuel in vehicle engines
for natural gas has been successful. Due to the low emission of toxic gases,
the use of biomethane in vehicles is considered one of the best renewable fuel
options in transportation [36, 37].
[38, 39]. In Sweden, it is estimated that more than 4,000 light vehicles and
fleets of public transport busses are moved by biomethane [32, 40]. On the
other hand, biodiesel and bioethanol are the biofuels most used in transport
vehicles in other countries. Biodiesel is a fuel formed by mono-alkyl esters of
long-chain fatty acids, produced from vegetable oils [41]. It can be used pure
(B100) or in diesel fuel mixtures, following the requirements of standards
allowed by the policies of each country (e.g., ASTM D6751) [5, 32, 42].
Author Copy
USA and Brazil are one of the main producing countries of biodiesel
in the world. In 2019, these countries produced around 6.5 and 5.9 billion
liters of biodiesel, behind Indonesia, responsible for the production of 7.9
billion liters [43] (Figure 17.1). Currently, the biodiesel industry in the USA
is the fastest-growing in the world, and it is expected to obtain favorable
investments and taxes in the coming years to encourage its expansion [33].
This biofuel is also widely used in the transport sector by the European
Union (EU), which in 2015 used approximately 12 billion liters [37, 44].
Among biofuels, bioethanol is the most promising to be used in the
worldwide transport sector in the short and medium-term (approximately
66% of the bioethanol produced in the world is used for transportation) [45,
46]. Bioethanol of first-generation is produced by a fermentative process of
feedstocks such as sugarcane, present in tropical and subtropical areas such
as in Brazil, India, and Colombia, or from corn, wheat, and tubers as occurs
in the USA, EU, and China [33, 47–50]. The bioethanol obtained from
sugarcane has approximately 70% lower GHG emissions than fuels formed
by conventional hydrocarbons, considering the life cycle balance [46].
In 2019, the USA produced the highest quantity of bioethanol in the
world, generating 59.73 billion of liters, followed by Brazil (32.48 billion
liters), EU (5.45 billion liters), China (3.41 billion liters), Canada (1.89
billion liters) (Figure 17.1) [51]. Although the USA is the largest producer
of bioethanol in the world, it still uses gasoline as its main transport fuel.
In 2019, petroleum fuels such as gasoline and diesel accounted for about
91% of the total energy use of the transportation sector in the USA [52]. In
this country, it is common to find a mixture of gasoline with 10% and 15%
bioethanol (E10–E15), the exact amount added varying by region [49, 53].
In this context, the current way of using biofuels is as additives for
gasoline and diesel, commonly without requirements of modification in the
vehicle engines. In its composition, bioethanol contains oxygen and using
mixtures E10 and E85, compared with fossil gasoline. Fuels combined with
bioethanol performed better compared to gasoline, in terms of depletion of
fossil fuels, with E10 and E85 were 6% and 64% to 70% lower, respectively.
To increase the use of bioethanol as an automotive biofuel, industries have
been working on the development of vehicles with a Flex system. In Brazil,
since 2003, automobiles with Flex engines have been sold, which have a
dual supply system for fuel, bioethanol, and/or gasoline. Those vehicles are
Author Copy
capable of using 100% bioethanol (E100), or gasoline, or any mixing rate
between them [29, 57]. Vehicles with this technology have become widely
used in Brazil, where about 95% of new cars sold have a Flex system [46, 58].
Although biofuels have been produced in many countries, the rate of
their adoption is still low. According to information from the International
Sustainable Development Strategy Group, in order to reduce the emission of
GHGs and fight against climate change, it is necessary to have sustainable
growth in the production of biofuel of 10% per year until 2030. Thus, political
support and innovation are needed to favor the use of available feedstocks,
to reduce industrial production costs, and to increase the advanced consump-
tion of biofuels [2, 189]. The current environmental impact caused by the
use of fossil fuels and the production of biofuels for the near future will be
discussed below.
The strong economic development is the direct reason for the increasing use
of fossil fuels, which causes the increase in global warming (GW) of the
planet and the harmful effects of natural destruction, such as the accelerated
melting of glaciers, floods, and long-term rains [34, 59, 60]. In addition, the
concentration of CO2 in the Earth’s atmosphere has grown 1.48 times from
280 ppm in 1,750 to 415 ppm in 2019, representing an increase of 0.9% per
year [34, 61].
To decarbonize the energy sector, several strategies are under development.
One of the established strategies is the utilization of renewable energy, which
production is expected to increase by 24% in the next decade [62]. However,
the dynamics of production of renewable energy and CO2 emissions in the
Author Copy
Challenges and Perspectives on Application of Biofuels 469
world may vary according to economic, political, or social global events [63].
For example, the CO2 emissions slowdown in 2019 was of 0.5%, but it would
need to be analyzed in the context of the strong increase in carbon emissions
(2.1%) due to the energy production boost in 2018 led especially by Russia,
Apple Academic Press
India, and USA [64]. Another issue that can be considered when analyzing
future trends would be the pandemic COVID-19 crisis, which occurred in
2020. This health crisis brought momentary environmental benefits, such
as the drastic decrease of global carbon emissions (expected to be 8%), but
also challenges as a decrease in global investment and energy demand (6%).
These factors together with economic issues, such as land distribution and
higher food prices, can also influence the production of biofuels [65, 66]. It
Author Copy
has been estimated that in the first period of 2020, global use of renewable
energy and electricity increased about 1.5% and 2%, respectively, compared
to the same period in 2019. However, for the first time in decades, it is
expected to decrease during this year [67].
It is important to highlight that modern renewable energy share has
increased only around 10% of the final global consumption of energy for
the last 10 years [3]. Thus, to achieve our global goals for the use of cleaner
energies and environmental improvement, the efforts must still focus, more
than ever, to improve the production and use of renewable energies as biofuels.
In the world, the biggest contributors to global energy growth in 2019 were
China, followed by Indonesia and India. China led the expansion of growth
and energy consumption, accounted for about 75% of total growth, and
achieved a greater increase in demand for each energy source [64]. Recent
reports show that with the growth of renewable energies in the world, fossil
fuel consumption fell for the first time to its lowest level. At the same time,
an increase of more than 60% in the global production of biofuels has been
observed since 2010, with a forecast to reach a consumption of 298 million
tons of oil equivalent (Mtoe) by 2030 [2, 3, 64]. Countries with the highest
production of biofuels are expected to double or triple their consumption of
biofuels in the next 10 years compared to their current production [2].
Biofuels such as bioethanol, biodiesel, and others (biomethane and
biohydrogen) are the most promising renewable energies to be produced on
a large scale and for being applied in the transport sector. For this reason,
countries have been increasing their investments to improve the production
of these biofuels to address the economically, environmentally, and socially
is estimated to increase more than 1% during the next decade [68]. As seen
before, USA and Brazil account for about 90% of bioethanol production
and approximately 85% of the global production of liquid biofuels [33, 69].
Both countries have policies that have encouraged the production and use
of bioethanol for decades. Many of these initiatives motivated other devel-
oping countries of America like Mexico, Colombia, and Argentina, and also
Asia, to introduce smaller scale bioethanol programs that are so beneficial
Author Copy
and necessary to achieve the international goals of sustainability [49]. As a
consequence, Argentina (forecasted total production growth of more than
46% by 2030) and other Asian countries (forecasted total production growth
between 19% and 50% by 2030) become large bioethanol producers. Nowa-
days, developed countries produce 58% of bioethanol in the world, and they
could increase their production by 6% until 2028 [68]. Other important top
producers are China (actual production of 10 billion liters), India (actual
production of 2 billion liters), and the EU (actual production of almost 7
billion liters), as is shown in Table 17.1.
Another liquid biofuel with positive impacts on the environment is
biodiesel. Nowadays, developed countries are responsible for 57% of the
world’s production of biodiesel, and it is expected by 2028 their produc-
tion should increase by 9% [68, 69]. Table 17.1 shows actualized data of
production and use of biodiesel around the world, having as greater producer
USA, followed by the EU. In addition, it was calculated that global biodiesel
production and consumption would increase by around 2% until 2030 [68, 69].
However, some of the developed countries do not produce and/or commer-
cialize biofuels, thus they do not appear in future production statistics.
Additionally, reports show that biogas is another potential biofuel, for
example, Dos Santos et al. [189], assessed energy potential and reduced
emissions by applying biogas energy formed from the biodigestion of
different organic wastes. It was found that potential power was between 4.5
and 6.9 GW, which would reduce CO2 emissions at a rate of 4.93% per year.
It is worth to note there are some specifications for each biogas applica-
tion or use. For example, for heat and electricity production usage biogas
requires the removal of water vapor and H2S. Its use in transportation, on
the other hand, requires an upgrading process that consists of impurities
removal (mainly CO2), conserving high energy content to reach each coun-
try’s quality standards [39, 70–72]. Fuels produced from biogas, such as
Global Total 127,584 126,302 143,111 12.1 –0.51 – 40,251 43,131 43,931 9.14 0.10 –
Global Consumption 125,864 127,250 143,190 13.7 – – 39,927 43,556 44,227 10.77 – –
World1 Emission – – 2.3j – – – – – 1.7j – – –
Saving (%)
Word Price (US Dollar 39.41 40.93 52.63 – – – 89.40 87.31 94.68 – – –
per Hectoliter)a
Overview of the annual and expected production, consumption, and price projections of bioethanol and biodiesel between the period 2018–2028,
Author Copy
472 Bioethanol: Biochemistry and Biotechnological Advances
liquefied biogas (LBG) and compressed natural gas (Bio-CNG) have been
commercially produced via biogas liquefaction and compression, respec-
tively [73]. Currently, there is a global production of around 3.5 Mtoe of
biomethane (obtained from biogas upgrading). Actually, biomethane repre-
Apple Academic Press
Author Copy
17.3.2 SUSTAINABILITY AND ENVIRONMENTAL BENEFITS OF
BIOFUELS
Even with government efforts for the constant increase of biofuels produc-
tion and promising future trends, there still are several challenges to be faced
to achieve the Sustainable Development Scenario (a major transformation
of the global energy systems, which is derived or aligned from other energy
scenarios). According to IRENA and IEA, the increase of approximately 3%
in the use of biofuels and renewables energies per year is a goal that would
allow environmental and energy sustainability in the next decades [2, 3]. For
the Sustainable Development Scenario, renewable energy share in the final
energy supply would increase by 17% by 2030 and 25% by 2050. There is
also a Transforming Energy Scenario that would bring sustainability in a
shorter time-lapse and reduction of total carbon emissions. To achieve this
goal, it would need an increase of six-fold and almost twice the actual trends
to achieve 28% of the renewable energy share in the final energy supply by
2030 and 66% by 2050 [3].
In the context of the Sustainable Development Scenario and taking
as an example the largest producers of biofuels and their forecast annual
production growth during the next decade, all these countries need at least
duplicate their predicted growth to achieve sustainable goals. For example,
in the case of USA, its actual production is of 35 Mtoe of biofuels represents
the 5% of its transportation energy sources [53], and have to increase to 95
Mtoe to get in the Sustainable Development Scenario, which means, this
country has to get its forecasted annual production of 1.9% to 7% of annual
production growth every year during the next decade. Other countries as
China and India would need to increase their actual production 8-fold to
fulfill the expected sustainable consumption of biofuels by 2030. According
to statistics, the EU would be the top producer with the hardest challenge, it
Author Copy
China 3 26 15.3 19
European 16 51 0.5 9
India 1 8 11.8 22
Source: International Energy Agency (IEA) [2]; Transport biofuels, IEA, Paris.
Author Copy
the restriction of biomass cultivation in areas intended for food planting [89].
Policies involving the production and use of biofuels have been under
development for many years in different countries. For example, Brazil
in 1975, developed the national fuel alcohol program (ProÁlcool), which
aimed the production of bioethanol from sugarcane, in response to the
oil crisis [90], while the National Biodiesel Production and Use Program
(PNPB) was launched in 2004 [191]. The development of these programs led
to the creation of the RenovaBio program, which integrated the production
of the most used biofuels (bioethanol, biodiesel, biogas), aiming at the
decarbonization of the Brazilian energy matrix and creating methodologies
that promote the stability of biofuels prices [91, 92].
China was another country that developed important programs for
biofuels production and application. In 1986, the 863 plan was created with
the National High Technology Research, which aimed to develop bioethanol
production from sweet potatoes and industrialize the biomass liquid fuel [8,
93, 94, 188]. Later in 2005, the Bioethanol Use Plan and the Renewable
Energy Law were created, which focused on the importance of biofuel,
promoting resources, and tax subsidies for its production [95, 96]. As part
of the bioethanol implementation plan, China has adopted different national
standards (GB 18350 and GB 18351) related to the production and use of
biofuels. A GB 18351 standard allows a mixture of 10% (E10) of ethanol with
gasoline [97]. In 2018, this mixture was already available in six provinces,
such as Liaoning, Heilongjiang, Henan, Jilin, Guangxi, and Anhui [98].
In 2006, the Ministry of Finance of China created financial support policies
for biofuels that encourage the development of biodiesel. These policies
provided subsidies for biodiesel plants that used non-food raw materials
[99–101]. The government strictly prohibited the use of used oil as cooking
oil, but illegal use persists due to the high profit [98]. The purchase price of
Author Copy
there is the standard renewable fuel program (RFS), which was established
by the energy policy act (EPA) in 2005, to increase the volume of renewable
fuel that is mixed with transport fuels. In 2007, this program was revised to
meet the requirements of the energy independence and security act (EISA),
that established specific targets for increase the production of cellulosic
biofuels, biomass diesel, and advanced biofuel in increasing volumes until
2022 [55, 65, 192].
In European Union (EU) countries, policies to encourage biofuels
application have evolved over the years. In 2008, was passed in the EU a
legislation that allowed the use of biofuels in the transport sector. One year
later, in 2009, as part of its “Climate Change Package,” the EU adopted
the Renewable Energy Directive (DRE 2009/28/EC). In this directive, each
Member State must achieve a minimum usage quota of 20% of energy from
renewable sources, reduce GHG emissions by 20% and increase energy
efficiency by 20% [104, 193]. In 2018, an agreement on the new Renewable
Energy Directive (REDII) was determined by the Council of the EU, which
will be in force between 2021 and 2030. This new directive includes the use
of advanced biofuels, and estimates consumption of total renewable energy
in the European energy matrix in 2030 should be 32% and in the transport
sector 14% [105, 106].
As seen above with the top biofuel producer countries, global govern-
ments have had initiatives to enhance the development of the biofuels
market. Among the most important initiatives created are:
1. Global Bioenergy Partnership (GBEP): Intergovernmental initia-
tive, with more than 70 members, including members of civil society,
government, and the private sector. It aims to support “the deploy-
ment of biomass and biofuels, particularly in developing countries
where the use of biomass is predominant” [107].
Author Copy
of low carbon technologies. New fuel production technologies are
starting to be deployed at new production facilities around the world.
All these projects aim to intensify the use of clean energy in the form of
biofuels; however, other limitations involve the economic barriers between
fossil and renewable fuels.
Even with fluctuations in the price of oil, biofuels still cost more than conven-
tional fossil fuels. The productive process of biofuels is the main barrier to
making these technologies economically viable. For this purpose, it is required
public support to encourage feedstock supply, cultivation of biomass, and
reduce the capital cost of biofuel processing [109, 110]. Moreover, the cost of
biofuel production depends mainly of process involved, industrial logistics,
obtaining biomass, storage, and distribution of the final product [111–114].
In the case of 1G and 2G biofuels the most important economic
consideration is related to the dependence on biomass [10]. The share in
cost of raw materials is 40% to 70% of the total production costs. Like
most of biofuels available, 1G biofuels depend on obtaining biomass,
involved in the food industries, which account for more than two-thirds
of total production costs [10]. Meanwhile, costs of 2G biofuels are also
dependent on additional steps (pretreatments) for the biomass involved
in the production processes [115]. For example, in the production of 2G
bioethanol, the highest costs are observed in the biomass conversion
stage [116, 193]. This step involves the use of enzymes for the hydrolysis
process, and these are responsible for about 20% to 40% of the total cost
of one liter of second-generation bioethanol [194].
Author Copy
FIGURE 17.2 Comparative cost analysis of main biofuels vs fossil fuels.
Source: [114, 117–121, 195].
Many of the political projects developed to date encourage the production and
increase availability of biofuels in society. Research carried out by special-
ized agencies, such as the International Energy Agency, shows that in 2030,
biofuels will contribute approximately 4% to 10% of the total energy involved
in the planet’s road transport. Thus, it will be necessary to use between 3.8%
and 4.5% of the arable land available worldwide for the cultivation of crops
directed to the production of biofuels, against only 1% of the total used [122].
According to the Global Renewable Energy Roadmap [30], the use
of biomass in several sectors in the world grows approximately 3.7% per
year between 2010 and 2030. It is estimated that in 2030 a total of 108 EJ
Author Copy
478 Mtoe of algae will be available worldwide, favoring the production of
2G and 3G biofuels [125, 196].
To get the expected demand for biofuel production in 2028, it will be
necessary to increase the availability of feedstocks involved in the production
process (Table 17.3). For example, in USA to obtain 59.72 ML of bioethanol
in 2019, about 374.02 kton of maize were used and to achieved the forecasted
amount for 2028, 388.70 kton will be required. Thus, countries that produce
and use biofuels must take action to ensure biomass availability, such as the
S2Biom Study, a project carried out in the EU that aims to provide sustainable
supplies (non-food biomass) to support the development of the bioeconomy [6].
Lignocellulosic biomass (LCB) stands out among the several available
feedstocks for the production of biofuels. This is an abundant organic resource
in many countries around the world, with the produce of an estimated 181.5
Gton annually [126–129]. Currently, about 8.2 Gton are used, of which
approximately 7 Gton of biomass are obtained from agricultural land, grass,
and forest, and 1.2 Gton are derived from agricultural waste [128]. During
2018/2019, Brazilian mills harvested around 620 Mton of sugarcane, gener-
ating about 180 Mton of wet bagasse and 43 Mton of cane straw [6, 197].
LCB represents a suitable low cost and eco-friendly alternative for the
production of biofuels and byproducts, however some aspects make difficult
the conversion of this material at large-scale that corresponds to its recalci-
trance, heterogeneity, and technical problems in its bioprocessing [130].
a
Thailand Sugarcane 28,600 68 40,183 Palm oil 3,410 50 5.215
Argentina Maize 49,157 9 73,450 Soybean oil 9,289 70 10.264
Global Consumption – 121.9 mln.L – – – 37.6 mln.L – –
as Biofuel Feedstock
Overview of the biomass used as feedstocks and production required of biomass for production of bioethanol and biodiesel between the period
2018–2028.
Author Copy
480 Bioethanol: Biochemistry and Biotechnological Advances
this key stage would mean a future large-scale application of biofuels in the
transport sector since it impacts in the global cost of the process. Furthermore,
other benefits would be market competition with fossil fuels, enhancement
of technology for energy transition, and reduction of negative environmental
Apple Academic Press
Author Copy
the high carbon and nitrogen ratio (C:N) of LCB negatively affects methane
production, and the mono-digestion of this substrate is not an adequate
strategy [126, 127, 129]. There are several methods to improve the use of
lignocellulosic feedstock as substrate in the AD process, such as anaerobic
co-digestion, bioaugmentation, and nutrient supplementation established
to optimize the C:N to 20–30:1 [132]. For example, Kainthola et al. [133]
evaluated the methane production of the anaerobic co-digestion of rice straw
and food waste. It was shown that methane yield increased around 94%
with respect to control in the optimized process (C:N 30, pH 7.32, food/
microorganism ratio 1.87).
In biodiesel production, LCB can be hydrolyzed and used for lipids
synthesis by oleaginous yeast, such as Meyerozyma guilliermondii and
Pichia kudriavzevii or filamentous fungus-like Mortierella isabellina for
subsequent obtaining of biodiesel [19, 134, 135]. Those microorganisms
could accumulate up to 60–80% lipids of their cell dry weight and can grow
exponentially while utilizing cheap substrates, becoming suitable candi-
dates for biotechnological purposes [136, 137]. The use of cheap carbon
sources like crop byproducts, significantly diminishes the cost of yeast oil
production [109].
Due to the complexity of the lignocellulosic matrix, the crystallinity of the
cellulose and the inhibition of metabolic operations by the lignin by-products,
other strategies have been established to improve the use of LCB in the
production of biofuels. Thus, pretreatments aim to modify the structural and
compositional barriers of this biomass, in order to improve digestibility by
enzymes to provide fermentable sugars for microorganisms [132, 138–140].
LCB pretreatment methods are widely known and can be classified in
physical, chemical, physicochemical, and biological processes, and can be
used alone or in combination.
Author Copy
costs in the overall process of bioenergy production from lignocellulose. For
example, Moiceanu et al. [144] used Miscanthus biomass in a 10 mm orifice
sieve, which means energy consumption of 50–67 kJ/kg, that led to the
increase of processing costs [144]. In addition, the energy consumption and
susceptibility of milling machines in the presence of large materials such as
stones and metals mixed with LCB, are some inconveniences [145]. Milling
is also a currently used pretreatment on biodiesel production, but its purpose
is centered on the release of lipids from marine algae species. A study with
Chlorella vulgaris shows that from 208 mg/L of biomass could be obtained
a lipid release of 10% when using bed milling [146].
On the other hand, chemical pretreatments like acid, alkaline, oxidative,
and organo-solvent processes have also been extensively studied for
pretreatment of LCB. Acid pretreatment causes the rupture of chemical bonds
present in the biomass matrix, mainly removing the hemicellulose fraction
[132]. One example of this pretreatment was shown by Taherdanak et al.
[147]. In this study was observed the decrease of the crystallinity index of the
wheat biomass accompanied by the solubilization of surface layers’ structure.
Moreover, Amnuaycheewa et al. [148] showed that biogas production and
saccharification from rice straw can be enhanced by pretreatment with organic
acids like oxalic and citric. Also, Huang et al. [149] used this pretreatment
to obtain sulfuric acid-treated rice straw hydrolysate (SARSH) for microbial
(Trichosporon fermentans) oil synthesis with potential for the production of
biodiesel. During fermentation using SARSH without detoxifying, 1.7 g/L
of lipid was obtained, a low yield compared to that obtained by fermentation
with detoxified hydrolyzate, generating a 40.1% lipid content. Despite of the
acid process is an efficient method for LCB, the main disadvantages are the
high cost of acids, drastic operation conditions (pressure and temperature),
and the formation of potential biological inhibitors [127, 150].
Author Copy
high temperatures and pressures in addition to intense shear [159]. Conse-
quently, highly reactive oxygen species (ROS) named hydroxyl radicals (OH)
are formed which produce oxidative degradation of LCB, turning more suit-
able for subsequent enzyme action and for microbial metabolization, which
could result in higher yields of biofuels [158, 160–163].
more than four months lit the warning signal for the 2020 sugarcane
harvest. Sugarcane has an annual cycle and, after its harvest, depends
on a regular water regime for the full development of the plant until
the next harvest [165]. As a consequence, this fact affected feedstock
Apple Academic Press
Author Copy
The problems reported related to 2G bioethanol production are based on
global industrial experiences. Industries such as Raízen, Granbio, Poet, Beta
Renewables, Abengoa (Synata Bio), and DuPont have found difficulties in
the production phase regarding the pretreatment processes, microorganisms
used, even in the mechanics and equipment costs of the unit operations
involved in the process [168].
Substantially, the production of 2G bioethanol is based on the use of
lignocellulosic and starchy biomass, which main challenges are pretreat-
ment and hydrolysis processes that are magnified on an industrial scale.
According to specialists working in 2G bioethanol plants, none of the
pretreatments and hydrolytic processes available today are more advanta-
geous than the other, and each varies in the production process, depending
on the type of raw material and several other factors [168]. For these stages,
enzymatic processes are preferred since they do not form microbial inhibitor
compounds, such as furans and phenolics. However, the enzymatic cocktails
used can make the process more expensive, due to their prices and the high
enzymatic load required. Thus, it is essential to improve the efficiency of
enzymes for reducing the price of this biofuel [168]. As stated by Schubert
[169], cellulase cocktails cost around 15–20 cents per gallon bioethanol as
compared to only 2–4 cents per gallon bioethanol for amylases used in the
bioethanol production process by starch (1G bioethanol). The adoption of
each type of enzyme in the pretreatment and enzymatic hydrolysis stages is
totally dependent on the structure of feedstocks.
It is well documented that the high crystallinity of cellulose, high levels
of lignin, hemicellulose, acetyl group, and degree of polymerization (DP) are
undesirable characteristics of biomass that directly interfere in the perfor-
mance of different enzyme cocktails [19, 170–175]. In an attempt to solve
such problems, companies such as Novozymes and Genencor International
harzianum. With the site-directed mutation, the active site of the engineered
enzyme became smaller and similar in size to that of β-glycosidases GH1.
The results of the analyzes indicated that the modified enzyme showed a
catalytic efficiency 300% higher than that of wild protein and became more
tolerant of glucose, providing a significant increase in the release of sugar
from all sources of plant biomass tested and also increase of thermal stability
during fermentation. This engineered β-glycosidase can be used to supple-
Author Copy
ment the enzymatic cocktails commercialized today for the degradation of
biomass and the production of second-generation biofuels. Besides, this
study shows that this enzyme can be produced by a homologous expression,
which makes possible its production on a large scale and with low costs.
In the case of the fermentation stage, the main challenge is the selection
of an appropriate microorganism to achieve the expected yield of bioethanol
production on an industrial scale. There are microorganisms more efficient
in the metabolism of hexoses from cellulose, than in pentoses from hemicel-
lulose hydrolysis [177]. In this context, there are several studies to select
wild yeast and produce engineered microorganisms also capable of ferment
pentoses.
Furthermore, other important difficulties are operational and mechanical
processes. GranBio reported the corrosion of the equipment in one of its
plants that used the steam explosion (SE) pretreatment in which the LCB
was subjected to very high temperature and pressure conditions. When the
system was suddenly decompressed, the biomass hit the walls of the equip-
ment causing routine damage and possibly interrupting the production. The
solution to this problem was the use of the liquid hot water (LHW) process
associated with the mechanical treatment of the fibers [177].
Raízen plant shows other problems related to the high level of dirt in the
biomass, which required a pre-cleaning step. It was also necessary to redesign
the process for the separation of sugars and lignin, introducing more filters
and centrifuges. In this 2G bioethanol plant, the following problems were also
presented: (I) the corrosion of pretreatment equipment, due to the silica present
in the sugarcane straw, (II) the excessive use of water when working with wheat
straw as feedstock, (III) the high water absorption by the sugarcane bagasse,
making it a high viscosity material, clogging of pipes and valves [177].
Also, according to Lynd [178], the success of the industrial implementa-
tion of 2G bioethanol depends on investments in research and development.
Author Copy
pretreated by mechanical or enzymatic methods to break cell walls, making
carbohydrates available. After this step, yeasts are added to convert ferment-
able sugars into bioethanol [180–182].
Some issues in the production of 3G bioethanol are reported as follows.
The selection of the microalgae strain is one of the most important
parameters. The carbohydrate composition of the cells must be considered
since the glycosidic portion acts as a substrate for the fermentation process.
As reported by Ueno [183], Chlorella vulgaris species is commonly used due
to its capacity of carbohydrate accumulation higher than 65%. According to
Kose and Oncel [184], the main parameters should be considerers for micro-
algae selection including (I) tolerance to oxygen gas levels, (II) resistance to
photo-inhibition, (III) high rates of cell growth, (IV) low need for nutrients
to develop, (V) high photosynthetic activity, (VI) easy adaptation to harsh
environmental conditions (for open systems), (VII) high survival rates, and
(VIII) genetic stability.
Temperature of microalgae cultivation is also an important factor that
influences the production of 3G bioethanol. As reported by Costa and De
Morais [180], it was observed that when Chlorella sp. was cultivated at
30°C, showed higher yields (448.0 μmol.g–1) of dry mass, in comparison
with cultures grown under 20°C (196.0 μmol.g–1). It is emphasized that the
drastic reduction in yield was also observed when the culture was performed
at 35°C and 45°C.
The microalgae cultivation mode also influences 3G bioethanol production.
Despite open lagoons are popular for being simple projects, contamination,
temperature control, sun exposure, and other factors, cannot be controlled,
highlighting the need for a careful selection of the bioreactor type [184].
Supply and physical and chemical characteristics of water are of funda-
mental importance for the seaweed cultivation during bioethanol 3G produc-
tion. A large volume of water is required to produce algae. Thus, this resource
can be optimized using wastewater, brackish, and with high levels of organic
nutrients [184].
Apple Academic Press
Biofuels are the main alternative to reduce environmental impact and cope
with the energy crisis, and their global production has significantly increased
in the last decade. In agreement with global sustainable goals, it is expected
to achieve a biofuels production growth by about 10% by 2030, using 28%
of available biomass as feedstock. Thus, it is extremely important to continue
Author Copy
investing and encouraging the development of technologies to make biofuels
an achievable option with a better cost-benefit for those countries without
oil reserves and refineries. Policies and programs, like RenovaBio in Brazil,
have been established to improve biofuel production and application in the
transportation sector. Consequently, biofuels developing countries could
support their future policies based on top producers’ experiences. Despite the
large number of biofuels manufactured in some countries, the actual industrial
production does not reach sustainable goals and still faces environmental,
economic, and technological challenges. At this moment, several strategies
for biofuel technologies need to be applied by the integration of different
biorefinery modes, such as 1G-1G integration (e.g., sugarcane-corn), 1.5G
(e.g., 1G-2G), and other possibilities based on 2G and 3G. In this context,
technological advances will be the core to boost biofuels to achieve a secure
transition from fossil fuels to an energy sustainable panorama.
KEYWORDS
• biofuels
• biomass conversion
• European union
• global warming
• greenhouse gas
• life cycle assessment
• technological challenges
• transport sector
REFERENCES
1. Rajagopal, D., & Zilberman, D., (2007). Review of Environmental, Economic and Policy
Aspects of Biofuels. The World Bank.
Apple Academic Press
Author Copy
& Water Magazine, 10, 9–21.
5. Beschkov, V., (2017). Biogas, biodiesel and bioethanol as multifunctional renewable fuels
and raw materials. In: JacobLopes, E., & Zepka, L. Q., (eds.), Frontiers in Bioenergy and
Biofuels (pp. 185–205).
6. International Energy Agency (IEA), (2020d). Advanced Biofuels–Potential for Cost
Reduction. Retrieved from: https://www.ieabioenergy.com/publications/new-publication-
advanced-biofuels-potential-for-cost-reduction/ (accessed on 28 October 2021).
7. Eisentraut, A., (2010). Sustainable Production of Second-generation Biofuels: Potential
and Perspectives in Major Economies and Developing Countries, 2010/01, 9–221.
Retrieved from: https://doi.org/10.1787/5kmh3njpt6r0-en.
8. Achinas, S., Horjus, J., Achinas, V., & Euverink, G. J. W., (2019). A pestle analysis of
biofuels energy industry in Europe. Sustainability, 11(21), 5981.
9. Demirbas, A., (2011a). Competitive liquid biofuels from biomass. Applied Energy,
88(1), 17–28.
10. Azad, A. K., Rasul, M. G., Khan, M. M. K., Sharma, S. C., & Hazrat, M. A., (2015).
Prospect of biofuels as an alternative transport fuel in Australia. Renewable and Sustainable
Energy Reviews, 43, 331–351.
11. Ahlgren, E. O., Börjesson, H. M., & Grahn, M., (2017). Transport biofuels in global
energy-economy modelling-a review of comprehensive energy systems assessment
approaches. GCB Bioenergy, 9(7), 1168–1180.
12. Azad, A. K., & Ameer, U. S. M., (2013). Performance study of a diesel engine by first-
generation bio-fuel blends with fossil fuel: An experimental study. Journal of Renewable
and Sustainable Energy, 5(1), 013118.
13. Lin, T., Rodríguez, L. F., Shastri, Y. N., Hansen, A. C., & Ting, K. C., (2014).
Integrated strategic and tactical biomass-biofuel supply chain optimization. Bioresource
Technology, 156, 256–266.
14. Demirbas, M. F., (2009b). Biorefineries for biofuel upgrading: A critical review. Applied
Energy, 86, S151–S161.
15. Naik, S. N., Goud, V. V., Rout, P. K., & Dalai, A. K., (2010). Production of first and
second-generation biofuels: A comprehensive review. Renewable and Sustainable
Energy Reviews, 14(2), 578–597.
16. Lee, R. A., & Lavoie, J. M., (2013). From first-to third-generation biofuels: Challenges
of producing a commodity from a biomass of increasing complexity. Animal Frontiers,
3(2), 6–11.
17. Lü, J., Sheahan, C., & Fu, P., (2011). Metabolic engineering of algae for fourth-generation
biofuels production. Energy & Environmental Science, 4(7), 2451–2466.
18. Abomohra, A. E. F., & Elshobary, M., (2019). Biodiesel, bioethanol, and biobutanol
production from microalgae. In: Microalgae Biotechnology for Development of Biofuel
Apple Academic Press
Author Copy
wood; biodiesel production using soybean. In: Biofuels, Solar and Wind as Renewable
Energy Systems (pp. 373–394). Springer, Dordrecht.
22. Demirbas, M. F., (2011b). Biofuels from algae for sustainable development. Applied
Energy, 88(10), 3473–3480.
23. Azad, A. K., Rasul, M. G., Khan, M. M. K., & Sharma, S. C., (2014). Review of biodiesel
production from microalgae: A novel source of green energy. In: Int. Green Energy Conf.
(pp. 1–9).
24. OECD Library, (2019). OECD Library-Enhancing Climate Change Migation through
Agriculture. OECD Organization for Economic Cooperation and Development
(database). Retrieved from: https://www.oecd-ilibrary.org/sites/dce06785-en/index.html?
itemId=/content/component/dce06785-en (accessed on 28 October 2021).
25. IEA, (2018). Renewable-2018: Analysis and Forecast to 2023, International Energy
Agency, Paris. International Energy Agency. Biofuels. Retrieved from: http://www.iea.
org/topics/biofuels/ (accessed on 28 October 2021).
26. International Energy Agency (IEA), (2019a). Renewables 2019. Retrieved from: https://
www.iea.org/reports/renewables-2019/transport#abstract (accessed on 28 October 2021).
27. OECD, (2018). Relatórios Econômicos OCDE Brasil, 2018. Retrieved from https://
epge.fgv.br/conferencias/apresentacao-do-relatorio-da-ocde-2018/files/relatorios-
economicos-ocde-brasil-2018.pdf (accessed on 28 October 2021).
28. International Energy Agency (IEA), (2019b). World Energy Model. Retrieved from:
Retrieved from: https://www.iea.org/reports/world-energy-model/sustainable-develop-
ment-scenario (accessed on 28 October 2021).
29. Nogueira, L. A. H., Souza, G. M., Cortez, L. A. B., & De Brito, C. C. H., (2020). Biofuels
for transport. In: Future Energy (pp. 173–197). Elsevier.
30. International Renewable energy Agency (IRENA), (2014). Global Bioenergy Supply
and Demand Projections: A Working Paper for REmap 2030, 1, 1–88.
31. Cremiato, R., Mastellone, M. L., Tagliaferri, C., Zaccariello, L., & Lettieri, P., (2018).
Environmental impact of municipal solid waste management using life cycle assessment:
The effect of anaerobic digestion, materials recovery and secondary fuels production.
Renewable Energy, 124, 180–188.
32. Callegari, A., Bolognesi, S., Cecconet, D., & Capodaglio, A. G., (2020). Production
technologies, current role, and future prospects of biofuels feedstocks: A state-of-the-art
review. Critical Reviews in Environmental Science and Technology, 50(4), 384–436.
33. Panchuk, M., Kryshtopa, S., Sładkowski, A., & Panchuk, A., (2020). Environmental
aspects of the production and use of biofuels in transport. In: Ecology in Transport:
Problems and Solutions (pp. 115–168). Springer, Cham. Retrieved from: https://link.
springer.com/chapter/10.1007/978-3-030-42323-0_3 (accessed on 28 October 2021).
Apple Academic Press
34. Sładkowski,A., (2020). Ecology in Transport: Problems and Solutions. Springer. Retrieved
from: https://link.springer.com/book/10.1007%2F978-3-030-42323-0 (accessed on 28
October 2021).
35. Statista, (2020a). Production of Biogas Worldwide in 2017, by Region. Retrieved from:
https://www.statista.com/statistics/481828/biogas-production-worldwide-by-region/
(accessed on 28 October 2021).
36. Scarlat, N., Dallemand, J. F., & Fahl, F., (2018). Biogas: Developments and perspectives
in Europe. Renewable Energy, 129, 457–472.
37. Eurostat, S. A., (2017). Your Key to European Statistics. Retrieved from: https://
Author Copy
ec.europa.eu/eurostat (accessed on 8 December 2021).
38. Lampinen, A., (2013). Development of biogas technology systems for transport.
Tekniikan Waiheita, 31(3), 5–37.
39. International Renewable Energy Agency (IRENA), (2018). Biogas for Road Vehicles:
Technology Brief. International Renewable Energy Agency, Abu Dhabi. https://www.
irena.org/-/media/Files/IRENA/Agency/Publication/2017/Mar/IRENA_Biogas_for_
Road_Vehicles_2017.pdf (accessed on 28 October 2021).
40. Van, F. F., (2012). Perspectives for Biogas in Europe. Oxford Institute for Energy Studies.
41. Gebremariam, S. N., & Marchetti, J. M., (2018). Biodiesel production economics.
Energy Conversion and Management,168, 74–84.
42. Burton, R., & Biofuels, P., (2008). An overview of ASTM D6751: Biodiesel standards
and testing methods. Alternative Fuels Consortium.
43. Statista, (2020b). Leading Biodiesel Producers Worldwide in 2019, by Country (in billion
liters). Retrieved from: https://www.statista.com/statistics/271472/biodiesel-production-
in-selected-countries/ (accessed on 28 October 2021).
44. European Commission Progress Reports, (2015). Retrieved from: https://ec.europa.
eu/neighbourhood-enlargement/system/files/2018–12/20151110_report_turkey.pdf
(accessed on 8 December 2021).
45. Mahapatra, M. K., & Kumar, A., (2017). A short review on biobutanol, a second-generation
biofuel production from lignocellulosic biomass. J. Clean Energy Technol., 5(1), 27–30.
46. Yun, Y., (2020). Alcohol fuels: Current status and future direction. In: Alcohol Fuels-
Current Technologies and Future Prospect. IntechOpen.
47. Cheng, J. J., & Timilsina, G. R., (2011). Status and barriers of advanced biofuel
technologies: A review. Renewable Energy, 36(12), 3541–3549.
48. Cardona, C. A., & Sánchez, Ó. J., (2007). Fuel ethanol production: Process design trends
and integration opportunities. Bioresource Technology, 98(12), 2415–2457.
49. Cortez, L. A., Griffin, M. W., Scaramucci, J. A., Scandiffio, M. I., & Braunbeck, O.
A., (2003). Considerations on the worldwide use of bioethanol as a contribution for
sustainability. Management of Environmental Quality: An International Journal.
50. Aydemir, E., Demirci, S., Dogan, A., Aytekin, A. Ö., & Sahin, F., (2014). Genetic
modifications of Saccharomyces cerevisiae for ethanol production from starch
fermentation: A review. Journal of Bioprocessing & Biotechniques, 4(7), 1.
51. Statista, (2020c). World production of fuel ethanol in 2019, by country (in millions of
gallons). Retrieved from: https://www.statista.com/statistics/281606/ethanol-production-
in-selected-countries/ (accessed on 28 October 2021).
Author Copy
case study. Journal of Cleaner Production, 245, 118933.
57. Cruz, C. H. B., Souza, G. M., & Cortez, L. A. B., (2014). Biofuels for transport. In:
Future Energy (pp. 215–244). Elsevier.
58. Belincanta, J., Alchorne, J. A., & Teixeira Da, S. M., (2016). The Brazilian experience
with ethanol fuel: Aspects of production, use, quality and distribution logistics. Brazilian
Journal of Chemical Engineering, 33(4), 1091–1102.
59. Karmaker, A. K., Rahman, M. M., Hossain, M. A., & Ahmed, M. R., (2020). Exploration
and corrective measures of greenhouse gas emission from fossil fuel power stations for
Bangladesh. Journal of Cleaner Production, 244, 118645.
60. Wang, S., Li, F., Wu, D., Zhang, P., Wang, H., Tao, X., Ye, J., & Nabi, M., (2018). Enzyme
pretreatment enhancing biogas yield from corn stover: Feasibility, optimization, and
mechanism analysis. Journal of Agricultural and Food Chemistry, 66(38), 10026–10032.
61. NOAA-National Oceanic and Atmospheric Administration (2020). Mauna Loa
Atmospheric Baseline Observatory. Retrieved from: https://www.noaa.gov/news/carbon-
dioxide-levels-in-atmosphere-hit-record-high-in-may (accessed on 28 October 2021).
62. C2ES, (2020). Center for climate and energy solution. Renewable Energy. https://www.
c2es.org/content/renewable-energy/#:~:text=Renewables%20made%20up%2017.1%20
percent,come%20from%20wind%20and%20solar (accessed on 28 October 2021).
63. World Bank, (2020). Global Economic Prospects. Washington, DC: World Bank.
64. BP, (2020). Statistical Review of World Energy. Retrieved from: https://www.bp.com/en/
global/corporate/energy-economics/statistical-review-of-world-energy.html (accessed on
28 October 2021).
65. Environmental Protection Agency (EPA), (n.d). Economics of Biofuels. Retrieved from:
https://www.epa.gov/environmental-economics/economics-biofuels#main-content
(accessed on 28 October 2021).
66. Anderson, W., (2019). Biofuels Effects: Social, Economic, Political & Environmental.
Schoolworkhelper Editorial Team. Retrieved from SchoolWorkHelper, https://school-
workhelper.net/biofuels-effect-on-social-economic-political-environmental/ (accessed
on 8 December 2021).
67. International Energy Agency (IEA), (2020b). The Impact of the Covid-19 Crisis on
Clean Energy Progress. Retrieved from https://www.iea.org/articles/the-impact-of-the-
covid-19-crisis-on-clean-energy-progress (accessed on 28 October 2021).
68. OECD/FAO, (2019a). OECD-FAO Agricultural Outlook 2019–2028, OECD Publishing,
Paris/Food and Agriculture Organization of the United Nations, Rome. Retrieved from
http://www.fao.org/3/CA4076EN/CA4076EN_Chapter9_Biofuels.pdf (accessed on 28
October 2021).
69. OECD/FAO, (2019b). OECD-FAO Agricultural Outlook. OECD Agriculture statistics
(database). doi: Dx.doi.org/10.1787/agr-outl-data-en http://www.fao.org/3/ca4076en/
Apple Academic Press
Author Copy
A. S., (2019). Recent updates on the production and upgrading of bio-crude oil from
microalgae. Bioresource Technology Reports, 7, 100216.
73. Yang, L., Ge, X., Wan, C., Yu, F., & Li, Y., (2014). Progress and perspectives in
converting biogas to transportation fuels. Renewable and Sustainable Energy Reviews,
40, 1133–1152.
74. International Energy Agency (IEA), (2020c). Outlook for Biogas and Biomethane:
Prospects for Organic Growth in World Energy Outlook Special Report. https://www.
iea.org/reports/outlook-for-biogas-and-biomethane-prospects-for-organic-growth/
an-introduction-to-biogas-and-biomethane (accessed on 28 October 2021).
75. Independent Statistics Analysis (EIA), (2020). July 2020 Monthly Energy Review US.
Retrieved from: https://www.eia.gov/totalenergy/data/monthly/pdf/mer.pdf (accessed on
28 October 2021).
76. United States Department of Agriculture-(USDA), (2019c). Brasil Biofuels Annual
2019. Retrieved from: https://apps.fas.usda.gov/newgainapi/api/report/downloadrepor
tbyfilename?filename=Biofuels%20Annual_Sao%20Paulo%20ATO_Brazil_8-9-2019.
pdf (accessed on 28 October 2021).
77. United States Department of Agriculture-(USDA). (2019d). Biofuels Annual, China
Will Miss E10 by 2020 Goal by Wide Margin. Retrieved from: https://apps.fas.usda.
gov/newgainapi/api/report/downloadreportbyfilename?filename=Biofuels%20
Annual_Beijing_China%20-%20Peoples%20Republic%20of_8-9-2019.pdf (accessed
on 28 October 2021).
78. United States Department of Agriculture-(USDA), (2019e). EU Biofuels Annual 2019.
Retrieved from: https://apps.fas.usda.gov/newgainapi/api/report/downloadreportbyfilena
me?filename=Biofuels%20Annual_The%20Hague_EU-28_7-15-2019.pdf (accessed on
28 October 2021).
79. United States Department of Agriculture-(USDA), (2019f). India Biofuels Annual 2019.
Retrieved from: https://apps.fas.usda.gov/newgainapi/api/report/downloadreportbyfilena
me?filename=Biofuels%20Annual_New%20Delhi_India_8-9-2019.pdf (accessed on 28
October 2021).
80. United States Department of Agriculture-(USDA), (2019g). Canada Biofuels Annual
2019, 1–613.
81. United States Department of Agriculture-(USDA), (2019h). Thailand Biofuels Annual
2019. Retrieved from: https://apps.fas.usda.gov/newgainapi/api/Report/DownloadRepor
tByFileName?fileName=Biofuels%20Annual_Bangkok_Thailand_11-04-2019 (accessed
on 28 October 2021).
Author Copy
86. Smith, M. N., (2016). The Number of Cars Worldwide is Set to Double by 2040. World
Economic Forum. Published online at weforum.org. Retrieved from: https://www.
weforum.org/agenda/2016/04/the-number-of-cars-worldwide-is-set-to-double-by-2040
(accessed on 28 October 2021).
87. Graver, B., Zhang, K., & Rutherford, D., (2019). Emissions from Commercial Aviation,
2018.
88. Pathak, L., & Shah, K., (2019). Renewable energy resources, policies and gaps in
BRICS countries and the global impact. Frontiers in Energy, 13(3), 506–521.
89. De Fátima, V. M., (2019). Northeastern Sugar and Ethanol Production, 129, 1–5.
90. Hoogwijk, M., Faaij, A., Van, D. B. R., Berndes, G., Gielen, D., & Turkenburg, W.,
(2003). Exploration of the ranges of the global potential of biomass for energy. Biomass
and Bioenergy, 25(2), 119–133.
91. Grassi, M. C. B., & Pereira, G. A. G., (2019). Energy-cane and RenovaBio: Brazilian
vectors to boost the development of Biofuels. Industrial Crops and Products, 129,
201–205.
92. Acharya, R. N., & Perez-Pena, R., (2020). Role of comparative advantage in biofuel
policy adoption in Latin America. Sustainability, 12(4), 1411.
93. Griliches, Z., (1957). Hybrid corn: An exploration in the economics of technological
change. Econometrica, Journal of the Econometric Society, 501–522.
94. Su, Y., Zhang, P., & Su, Y., (2015). An overview of biofuels policies and industrialization
in the major biofuel producing countries. Renewable and Sustainable Energy Reviews,
50, 991–1003.
95. Utkulu, U., & Seymen, D., (2004). Revealed comparative advantage and competitiveness:
Evidence for Turkey vis-à-vis the EU/15. In: European Trade Study Group 6th Annual
Conference, ETSG (pp. 1–26).
96. Lapan, H., & Moschini, G., (2012). Second-best biofuel policies and the welfare
effects of quantity mandates and subsidies. Journal of Environmental Economics and
Management, 63(2), 224–241.
97. Saravanan, A. P., Pugazhendhi, A., & Mathimani, T., (2020). A comprehensive
assessment of biofuel policies in the BRICS nations: Implementation, blending target
and gaps. Fuel, 272, 117635.
98. Hao, H., Liu, Z., Zhao, F., Ren, J., Chang, S., Rong, K., & Du, J., (2018). Biofuel for
vehicle use in China: Current status, future potential and policy implications. Renewable
and Sustainable Energy Reviews, 82, 645–653.
99. Qiu, H., Sun, L., Huang, J., & Rozelle, S., (2012). Liquid biofuels in China: Current status,
government policies, and future opportunities and challenges. Renewable and Sustainable
Energy Reviews, 16(5), 3095–3104.
100. Chang, S., Zhao, L., Timilsina, G. R., & Zhang, X., (2012). Biofuels development in China:
Apple Academic Press
Technology options and policies needed to meet the 2020 target. Energy Policy, 51, 64–79.
101. Lu, Y. H., & Yang, Y. Q., (2019). Sugarcane Biofuels Production in China. In: Sugarcane
Biofuels (pp. 139–155). Springer, Cham.
102. Schuman, S., & Lin, A., (2012). China’s Renewable Energy Law and its impact on
renewable power in China: Progress, challenges and recommendations for improving
implementation. Energy Policy, 51, 89–109.
103. Nahm, J., (2017). Exploiting the implementation gap: Policy divergence and industrial
upgrading in China’s wind and solar sectors. The China Quarterly, 231, 705–727.
104. European Commission, (2020). Biofuels. Retrieved from: https://ec.europa.eu/energy/
Author Copy
topics/renewable-energy/biofuels/overview_en (accessed on 28 October 2021).
105. EU, (2018). Council-Council of the European Union. Proposal for a directive of the Euro-
pean Parliament and of the council on the promotion of the use of energy from renewable
sources (recast). Brussels. Available at: http://data.consilium.europa.eu/ (accessed on 28
October 2021).
106. European Commission, (2019). Renewable Energy – Recast to 2030, (RED II). Retrieved
from: https://ec.europa.eu/jrc/en/jec/renewable-energy-recast-2030-red-ii (accessed on
28 October 2021).
107. Global bioenergy. (GBEP), (2016). Retrieved from: http://www.globalbioenergy.org/
fileadmin/user_upload/gbep/docs/ENGLISH_Background_note_GBEP_Setember_
2016_FINAL.pdf (accessed on 28 October 2021).
108. Biofuture Platform, (2020). Retrieved from: http://www.biofutureplatform.org/ (accessed
on 28 October 2021).
109. Demirbas, A., (2009c). Political, economic and environmental impacts of biofuels: A
review. Applied Energy, 86, S108–S117.
110. Demırbas, A., (2017). The social, economic, and environmental importance of biofuels
in the future. Energy Sources, Part B: Economics, Planning, and Policy, 12(1), 47–55.
111. Van, R. R., Sanders, J., Bakker, R., Blaauw, R., Zwart, R., & Van, D. D. B., (2011).
Biofuel-driven biorefineries for the co-production of transportation fuels and added-value
products. In: Handbook of Biofuels Production (pp. 559–580). Woodhead Publishing.
112. Cremonez, P. A., Feroldi, M., Feiden, A., Teleken, J. G., Gris, D. J., Dieter, J., &
Antonelli, J., (2015). Current scenario and prospects of use of liquid biofuels in South
America. Renewable and Sustainable Energy Reviews, 43, 352–362.
113. Lucena, A. F., Clarke, L., Schaeffer, R., Szklo, A., Rochedo, P. R., Nogueira, L. P., &
Kober, T., (2016). Climate policy scenarios in Brazil: A multi-model comparison for
energy. Energy Economics, 56, 564–574.
114. Guimarães, L. N., (2020). Is there a Latin American electricity transition? A snapshot
of intraregional differences. In: The Regulation and Policy of Latin American Energy
Transitions (pp. 3–20). Elsevier.
115. Srivastava, N., Kharwar, R. K., & Mishra, P. K., (2019). Cost economy analysis of
biomass-based biofuel production. In: New and Future Developments in Microbial
Biotechnology and Bioengineering (pp. 1–10). Elsevier.
116. Soratana, K., Harden, C. L., Zaimes, G. G., Rasutis, D., Antaya, C. L., Khanna, V., &
Landis, A. E., (2014). The role of sustainability and life cycle thinking in US biofuels
policies. Energy Policy, 75, 316–326.
117. Tozer, L., (2019). The urban material politics of decarbonization in Stockholm, London
and San Francisco. Geoforum, 102, 106–115.
118. Silva, N. D. L. D., (2010). Biodiesel production = process and characterizations, 1–192.
Retrieved from: http://www.repositorio.unicamp.br/handle/REPOSIP/266978.
Apple Academic Press
119. Lu, C., (2018). When will biofuels be economically feasible for commercial flights?
Considering the difference between environmental benefits and fuel purchase costs.
Journal of Cleaner Production, 181, 365–373.
120. Markel, E., Sims, C., & English, B. C., (2018). Policy uncertainty and the optimal invest-
ment decisions of second-generation biofuel producers. Energy Economics, 76, 89–100.
121. Li, B., Li, Y., Liu, H., Liu, F., Wang, Z., & Wang, J., (2017). Combustion and emission
characteristics of diesel engine fueled with biodiesel/PODE blends. Applied Energy, 206,
425–431.
122. Alisson, E., (2013). Biofuels Face Challenges for Expansion, 1–5. Retrieved from:
Author Copy
https://agencia.fapesp.br/biocombustiveis-enfrentam-desafios-para-expansao/18259/
(accessed on 28 October 2021).
123. Gerbens-Leenes, P. W., (2017). Bioenergy water footprints, comparing first, second and
third-generation feedstocks for bioenergy supply in 2040. European Water, 59, 373–380.
124. Khanna, M., Wang, W., Hudiburg, T. W., & DeLucia, E. H., (2017). The social inefficiency
of regulating indirect land-use change due to biofuels. Nature Communications, 8(1), 1–9.
125. Espi, E., Ribas, I., Diaz, C., & Sastron, O., (2020). Feedstocks for advanced biofuels. In:
Sustainable Mobility. IntechOpen.
126. Paul, S., & Dutta, A., (2018). Challenges and opportunities of lignocellulosic biomass
for anaerobic digestion. Resources, Conservation and Recycling, 130, 164–174.
127. Baruah, J., Nath, B. K., Sharma, R., Kumar, S., Deka, R. C., Baruah, D. C., & Kalita,
E., (2018). Recent trends in the pretreatment of lignocellulosic biomass for value-added
products. In: Frontiers in Energy Research (Vol. 6, p. 141).
128. Dahmen, N., Lewandowski, I., Zibek, S., & Weidtmann, A., (2019). Integrated
lignocellulosic value chains in a growing bioeconomy: Status quo and perspectives.
GCB Bioenergy, 11(1), 107–117.
129. Cheah, W. Y., Sankaran, R., Show, P. L., Ibrahim, T. N. B. T., Chew, K. W., Culaba,
A., & Jo-Shu, C., (2020). Pretreatment methods for lignocellulosic biofuels production:
Current advances, challenges and future prospects. Biofuel Research Journal, 7(1), 1115.
130. Sharma, R., Joshi, R., & Kumar, D., (2020). Present status and future prospect of genetic
and metabolic engineering for biofuels production from lignocellulosic biomass. In:
Genetic and Metabolic Engineering for Improved Biofuel Production from Lignocellulosic
Biomass (pp. 171–192). Elsevier.
131. Russel, T. H., & Frymier, P., (2012). Bioethanol production in Thailand: A teaching case
study comparing cassava and sugar cane molasses. The Journal of Sustainability Education.
132. Abraham, A., Mathew, A. K., Park, H., Choi, O., Sindhu, R., Parameswaran, B., & Sang,
B. I., (2020). Pretreatment strategies for enhanced biogas production from lignocellulosic
biomass. Bioresource Technology, 301, 122725.
133. Kainthola, J., Kalamdhad, A. S., & Goud, V. V., (2020). Optimization of process
parameters for accelerated methane yield from anaerobic co-digestion of rice straw and
food waste. Renewable Energy, 149, 1352–1359.
134. Ruan, Z., Zanotti, M., Wang, X., Ducey, C., & Liu, Y., (2012). Evaluation of lipid accu-
mulation from lignocellulosic sugars by Mortierella isabellina for biodiesel production.
Bioresource Technology, 110, 198–205.
135. Ananthi, V., Siva, P. G., Chang, S. W., Ravindran, B., Nguyen, D. D., Vo, D. V. N.,
La, D. D., et al., (2019). Enhanced microbial biodiesel production from lignocellulosic
hydrolysates using yeast isolates. Fuel, 256, 115932.
136. Patel, A., Arora, N., Sartaj, K., Pruthi, V., & Pruthi, P. A., (2016). Sustainable biodiesel
Apple Academic Press
Author Copy
139. Hendriks, A. T. W. M., & Zeeman, G., (2009). Pretreatments to enhance the digestibility
of lignocellulosic biomass. Bioresource Technology, 100(1), 10–18.
140. Dziekońska-Kubczak, U. A., Berłowska, J., Dziugan, P. T., Patelski, P., Balcerek, M.,
Pielech-Przybylska, K. J., & Domański, J. T., (2018). Comparison of steam explosion,
dilute acid, and alkali pretreatments on enzymatic saccharification and fermentation of
hardwood sawdust. BioResources, 13(3), 6970–6984.
141. Victorin, M., Davidsson, Å., & Wallberg, O., (2020). Characterization of mechanically
pretreated wheat straw for biogas production. Bioenergy Research, 1–12.
142. Dai, X., Hua, Y., Dai, L., & Cai, C., (2019). Particle size reduction of rice straw enhances
methane production under anaerobic digestion. Bioresource Technology, 293, 122043.
143. Dell,’O. P. P., & Spena, V. A., (2020). Mechanical pretreatment of lignocellulosic
biomass to improve biogas production: Comparison of results for giant reed and wheat
straw. Energy, 203, 117798.
144. Moiceanu, G., Paraschiv, G., Voicu, G., Dinca, M., Negoita, O., Chitoiu, M., & Tudor, P.,
(2019). Energy consumption at size reduction of lignocellulose biomass for bioenergy.
Sustainability, 11(9), 2477.
145. Dahunsi, S. O., (2019). Mechanical pretreatment of lignocelluloses for enhanced biogas
production: Methane yield prediction from biomass structural components. Bioresource
Technology, 280, 18–26.
146. Zheng, H., Yin, J., Gao, Z., Huang, H., Ji, X., & Dou, C., (2011). Disruption of Chlorella
vulgaris cells for the release of biodiesel-producing lipids: A comparison of grinding,
ultrasonication, bead milling, enzymatic lysis, and microwaves. Applied Biochemistry
and Biotechnology, 164(7), 1215–1224.
147. Taherdanak, M., Zilouei, H., & Karimi, K., (2016). The influence of dilute sulfuric acid
pretreatment on biogas production from wheat plant. International Journal of Green
Energy, 13(11), 1129–1134.
148. Amnuaycheewa, P., Hengaroonprasan, R., Rattanaporn, K., Kirdponpattara, S.,
Cheenkachorn, K., & Sriariyanun, M., (2016). Enhancing enzymatic hydrolysis and
biogas production from rice straw by pretreatment with organic acids. Industrial Crops
and Products, 87, 247–254.
149. Huang, C., Chen, X. F., Xiong, L., & Ma, L. L., (2012). Oil production by the yeast
Trichosporon dermatitis cultured in enzymatic hydrolysates of corncobs. Bioresource
Technology, 110, 711–714.
150. Jönsson, L. J., & Martín, C., (2016). Pretreatment of lignocellulose: Formation of inhibi-
tory by-products and strategies for minimizing their effects. In: Bioresource Technology
(Vol. 199, pp. 103–112). Elsevier Ltd.
151. Yilmaz, F., Kökdemir, Ü. E., & Perendeci, N. A., (2019). Enhancement of methane
Apple Academic Press
Author Copy
Technology, 41(8), 997–1006.
154. Bimestre, T. A., Júnior, J. A. M., Botura, C. A., Canettieri, E. V., & Tuna, C. E., (2020).
Theoretical modeling and experimental validation of hydrodynamic cavitation reactor
with a Venturi tube for sugarcane bagasse pretreatment. Bioresource Technology, 311,
123540.
155. Terán H. R., Kamoei, D. V., Ahmed, M. A., Da Silva, S. S., Han, J. I., & Santos, J.
C. D., (2018). A new approach for bioethanol production from sugarcane bagasse
using hydrodynamic cavitation assisted-pretreatment and column reactors. Ultrasonics
Sonochemistry, 43, 219–226.
156. Terán H. R., Dionízio, R. M., Sánchez, M. S., Prado, C. A., De Sousa, J. R., Da Silva, S.
S., & Santos, J. C., (2020). Hydrodynamic cavitation-assisted continuous pretreatment
of sugarcane bagasse for ethanol production: Effects of geometric parameters of the
cavitation device. Ultrasonics Sonochemistry, 63.
157. Nalawade, K., Saikia, P., Behera, S., Konde, K., & Patil, S., (2020). Assessment of
multiple pretreatment strategies for 2G L-lactic acid production from sugarcane bagasse.
Biomass Conversion and Biorefinery, 1–14.
158. Nagarajan, S., & Ranade, V. V., (2020). Pretreatment of distillery spent wash (vinasse)
with vortex based cavitation and its influence on biogas generation. Bioresource
Technology Reports, 100480.
159. Gogate, P. R., & Pandit, A. B., (2005). A review and assessment of hydrodynamic cavitation
as a technology for the future. Ultrasonics Sonochemistry, 12(1, 2 SPEC. ISS.), 21–27.
160. Zieliński, M., Dębowski, M., Kisielewska, M., Nowicka, A., Rokicka, M., & Szwarc,
K., (2019). Cavitation-based pretreatment strategies to enhance biogas production in a
small-scale agricultural biogas plant. Energy for Sustainable Development, 49, 21–26.
161. Terán H. R., Dos, S. J. C., Ahmed, M. A., Jeon, S. H., Da Silva, S. S., & Han, J. I.,
(2016). Hydrodynamic cavitation-assisted alkaline pretreatment as a new approach for
sugarcane bagasse biorefineries. Bioresource Technology, 214, 609–614.
162. Patil, P. N., Gogate, P. R., Csoka, L., Dregelyi-Kiss, A., & Horvath, M., (2016). Inten-
sification of biogas production using pretreatment based on hydrodynamic cavitation.
Ultrasonics Sonochemistry, 30, 79–86.
163. Huang, C., Zong, M. H., Wu, H., & Liu, Q. P., (2009). Microbial oil production from
rice straw hydrolysate by Trichosporon fermentans. Bioresource Technology, 100(19),
4535–4538.
164. Morrow, III. W. R., Gopal, A., Fitts, G., Lewis, S., Dale, L., & Masanet, E., (2014).
Feedstock loss from drought is a major economic risk for biofuel producers. Biomass
and Bioenergy, 69, 135–143.
165. Gomes, (2018). Drought lights up alert for Brazil's sugarcane crop in-2019. Retrieved
Apple Academic Press
from: https://www.novacana.com/n/cana/safra/seca-alerta-safra-cana-brasil-2019-200718
(accessed on 28 October 2021).
166. Pugliese, L., Lourencetti, C., & Ribeiro, M. L., (2017). Environmental impacts on
Brazilian ethanol production: A brief discussion from field to industry. Multidisciplinary
Brazilian Journal, 20(1), 142–165.
167. Costa, C. C. D., & Burnquist, H. L., (2016). Impacts of gasoline price controls on biofuel
ethanol in Brazil. Economic Studies (São Paulo), 46(4), 1003–1028.
168. Nova Cana, (2017). Dilemma of Ethanol 2G: The challenges of cellulosic feedstock.
Retrieved from https://www.novacana.com/n/etanol/2-geracao-celulose/dilema-etanol-
Author Copy
2g-desafios-materia-prima-celulosica-260117 (accessed on 28 October 2021).
169. Schubert, C., (2006). Can biofuels finally take center stage?. Nature Biotechnology,
24(7), 777–784.
170. Nazhad, M. M., Ramos, L. P., Paszner, L., & Saddler, J. N., (1995). Structural constraints
affecting the initial enzymatic hydrolysis of recycled paper. Enzyme and Microbial
Technology, 17(1), 68–74.
171. Chang, V. S., & Holtzapple, M. T., (2000). Fundamental factors affecting biomass
enzymatic reactivity. In: Twenty-First Symposium on Biotechnology for Fuels and
Chemicals (pp. 5–37). Humana Press, Totowa, NJ.
172. Zhu, J. Y., Wang, G. S., Pan, X. J., & Gleisner, R., (2008). The status of and key barriers
in lignocellulosic ethanol production: A technological perspective. In: International
Conference on Biomass Energy Technologies: Guangzhou, China, December 3–5, 2008
(Vol. 1, pp. 1–12). [Guangzhou, China]: Guangzhou Institute of Energy Conversion, The
Chinese Academy of Science.
173. Banerjee, S., Mudliar, S., Sen, R., Giri, B., Satpute, D., Chakrabarti, T., & Pandey, R.
A., (2010). Commercializing lignocellulosic bioethanol: Technology bottlenecks and
possible remedies. Biofuels, Bioproducts and Biorefining: Innovation for a Sustainable
Economy, 4(1), 77–93.
174. Chen, H., Han, Q., Venditti, R. A., & Jameel, H., (2015). Enzymatic hydrolysis of
pretreated newspaper having high lignin content for bioethanol production. BioResources,
10(3), 4077–4098.
175. Abo, B. O., Gao, M., Wang, Y., Wu, C., Ma, H., & Wang, Q., (2019). Lignocellulosic
biomass for bioethanol: An overview on pretreatment, hydrolysis and fermentation
processes. Reviews on Environmental Health, 34(1), 57–68.
176. Santos, C. A., Morais, M. A., Terrett, O. M., Lyczakowski, J. J., Zanphorlin, L. M.,
Ferreira-Filho, J. A., & Souza, A. P., (2019). An engineered GH1 β-glucosidase displays
enhanced glucose tolerance and increased sugar release from lignocellulosic materials.
Scientific Reports, 9(1), 1–10.
177. Marques, F., (2018). FAPESP Magazine: Obstacles on the way. Retrieved from: https://
revistapesquisa.fapesp.br/obstaculos-no-caminho/ (accessed on 28 October 2021).
178. Lynd, L. R., (2017). The grand challenge of cellulosic biofuels. Nature Biotechnology,
35(10), 912.
179. Özçimen, D., Koçer, A. T., İnan, B., & Özer, T., (2020). Bioethanol production from
microalgae. In: Handbook of Microalgae-Based Processes and Products (pp. 373–389).
Academic Press.
180. Costa, J. A. V., & De Morais, M. G., (2011). The role of biochemical engineering in the
production of biofuels from microalgae. Bioresource Technology, 102(1), 2–9.
181. Amin, S., (2009). Review on biofuel oil and gas production processes from microalgae.
Energy Conversion and Management, 50(7), 1834–1840.
Apple Academic Press
182. Kumar, N., Sonthalia, A., & Pali, H. S., (2020). Next-generation biofuels—opportunities
and challenges. In: Innovations in Sustainable Energy and Cleaner Environment (pp.
171–191). Springer, Singapore.
183. Ueno, Y., Kurano, N., & Miyachi, S., (1998). Ethanol production by dark fermentation
in the marine green alga, Chlorococcum littorale. Journal of Fermentation and Bioengi-
neering, 86(1), 38–43.
184. Kose, A., & Oncel, S. S., (2017). Algae as a promising resource for biofuel industry:
Facts and challenges. International Journal of Energy Research, 41(7), 924–951.
185. Dos, S. I. F. S., Vieira, N. D. B., De Nóbrega, L. G. B., Barros, R. M., & Tiago, F. G. L.,
Author Copy
(2018). Assessment of potential biogas production from multiple organic wastes in Brazil:
Impact on energy generation, use, and emissions abatement. Resources, Conservation
and Recycling, 131, 54–63.
186. NREL (Transforming Energy), (2018). Renewable Energy Data Book. Retrieved from
https://www.nrel.gov/docs/fy20osti/75284.pdf (accessed on 28 October 2021).
187. U.S. Energy Information Administration (USEIA). Total Petroleum and Other Liquids
Production. (2014). Retrieved from: https://www.eia.gov/forecasts/steo/reDort/global
oil.cfin.
188. Kumar, D., & Singh, V. (2019). Bioethanol production from corn. In Corn (pp. 615–631).
AACC International Press.
189. International Energy Agency (IEA). (2019c). IEA convenes 2019 meeting of the
Renewable Industry Advisory Board. Retrieved from: https://www.iea.org/news/iea-
convenes-2019-meeting-of-the-renewable-industry-advisory-board.
190. OECD/FAO (2020), OECD-FAO Agricultural Outlook 2020–2029, FAO, Rome/OECD
Publishing, Paris, https://doi.org/10.1787/1112c23b-en. https://www.fao.org/3/ca8861en/
CA8861EN.pdf.
191. Laursen, K. (2015). Revealed comparative advantage and the alternatives as measures of
international specialization. Eurasian Business Review, 5(1), 99–115.
192. Congressional Research Service (CRS). (2019). The Renewable Fuel Standard (RFS):
An Overview. Retrieved from: https://crsreports.congress.gov/product/pdf/R/R43325
193. Subramaniam, Y., Masron, T. A., & Azman, N. H. N. (2020). Biofuels, environmental
sustainability, and food security: A review of 51 countries. Energy Research & Social
Science, 68, 101549.
194. Centro Nacional de Pesquisa em Energia e Materiais. (CNPEM) (2017). Retrieved from:
http://pages.cnpem.br/2gbioethanol/wp-content/uploads/sites/85/2017/06/Jose_ Bressiani_
Granbio.pdf (accessed on 8 December 2021).
195. Carriquiry, M., Elobeid, A., Dumortier, J., & Goodrich, R. (2020). Incorporating
sub-national Brazilian agricultural production and land-use into US biofuel policy
evaluation. Applied Economic Perspectives and Policy, 42(3), 497–523.
196. International Energy Agency (IEA). (2019). Bioenergy. Annual Report 2018. Paris: IEA;
2019. Retrieved from: ieabioenergy.com/wp-content/uploads/2020/05/IEA-Bioenergy-
Annual-Report-2019.pdf (accessed on 8 December 2021).
197. CONAB (2019). Monitoring of the Brazilian crop: coffee. Brasilia: Conab. Retrieved
from: https://www.conab.gov.br (accessed on 8 December 2021).
Author Copy
32, 33, 46, 47, 55, 57, 58, 60, 63, 69, 71,
2-keto-3-deoxy-6-phosphogluconic acid 85, 104, 113, 123, 142, 145–147, 155,
(KDPG), 28 162–165, 172, 181, 195, 196, 198–200,
210–212, 219, 272, 274, 275, 279, 286,
3 325, 339, 391, 422, 425, 428, 474
3-hydroxyanthranilic acid (HAA), 429 acetyltransferases, 63
dehydrogenase (ADH), 6, 16, 24, 32, 46,
5 47, 55, 57, 58, 60, 69, 71, 85, 104, 113,
5-hydroxymethylfurfural (HMF), 59, 104 123, 155
ADH II (ADH2), 55, 57, 59, 68, 123, 145
A enzymes, 6, 86
Alcoholic
Abiotic agents, 75 beverage, 2, 14, 21–23, 143
Acetaldehyde, 6, 24, 31, 32, 47, 55, 57, 63, fermentation, 16, 24, 26, 30, 65, 197, 240,
75, 85, 104, 155 254, 432, 433
aldehyde dehydrogenase (ALDH), 155 Aldehyde, 6, 57, 58, 61, 63–65, 73, 76–78,
dehydrogenase, 47 117, 118, 181
Acetate kinase, 47, 121
dehydrogenase, 32, 58, 63, 69, 118, 155
Acetic acid, 45, 61–64, 70, 104, 105, 113,
gene, 32
115–117, 170, 181, 348, 356, 374, 380,
groups, 63
385
Algal biomass, 486
Acetyl-CoA reduction, 155
Alphaproteobacteria, 27
Acid
base treatment, 347 American civil war, 162
tolerance, 27 Amino acid, 7, 29, 30, 32, 57, 59, 67, 70,
tolerant mutants, 115 73, 76, 99, 103, 105, 118, 145, 147, 155,
Acidity, 1, 3, 33, 321 345
Acidophilic mesophiles, 10 Ammonia, 8, 180, 284, 339, 352, 353, 356,
Adaptation laboratory evolution, 105 409, 410, 413
Adhesives, 14 fiber expansion (AFEX), 180, 339, 340,
Aerobic 351, 352, 409
conditions, 6, 31, 35, 70, 76 recycling percolation (ARP), 180
gram-negative bacteria, 74 Ammonium
Agave reductase, 8
americana, 371, 378, 380 salts, 6, 8
hydrothermal pretreatment, 383 Amorphous macromolecule, 107
waste, 377, 379 Amyloglucosidase, 181, 278
Author Copy
Aquatic biomass, 373 329, 330, 339, 371, 406, 419, 446, 447,
Aqueous two-phase systems (ATPS), 222 464, 466, 471, 474, 479, 483
Arabinose, 73, 107, 115, 178, 236, 286, 380, challenges
424, 432, 433 first-generation bioethanol, 482
Aromatic second-generation bioethanol, 484
amino acids (AAA), 30, 149 third-generation bioethanol, 486
molecules, 65 commercialization, 161, 162
Artificial fermentation, 181, 391
antioxidant defense system, 106 production, 22, 109, 113, 116, 120, 161,
modification, 107 256, 329, 371, 406, 419, 446, 447, 483
Aspartic acids, 8 enzymatic hydrolysis, 238
Autohydrolysis, 375, 382, 383 fermentation, 240
Autophagy, 100 genetic engineering strategies, 108
Azeotrope, 196, 197, 199, 202, 203, 206 microorganisms, 25
Azeotropic pretreatment, 237
composition, 196–199, 203, 206, 207, process, 26, 236
211, 213 substrates, 25
distillation, 195, 196, 199–201, 210, 216, yield, 26
yield, 26, 33, 183, 243, 325, 327–329
282
Biofilm reactor, 36
Biofuel, 21–23, 26, 44, 48, 49, 56, 98–100,
B
102, 104, 105, 119, 123, 162, 165, 166,
Bacterial 170, 172, 173, 175, 234, 246, 272, 273, 283,
gene expression, 74 295, 324, 343, 377, 405–409, 412, 420,
pathway, 29 423, 425, 430–432, 446, 447, 463–467,
Batch, 14, 282, 451 469–471, 474–478, 481, 482, 484, 487
system, 14, 250, 403 challenges, 474
Beta vulgaris, 165, 166 production, 98, 102, 105, 108, 122, 123,
Binary permeability, 207 170, 176, 255, 316, 325, 341, 467, 469,
Biocatalysis, 99 472, 474, 487
Biocatalyst, 21, 143, 425, 451, 453 solvent tolerance, 102
Bio-char, 373 sustainability and environmental benefits,
Biochemical, 373, 374 472
degradation pathways, 24 Biofuture platform (BFP), 476
Biodegradability, 180 Biogas, 117, 282, 344, 372, 377, 405, 452,
Biodetoxification, 118 464, 465, 468, 470, 472, 474, 478, 480, 481
Author Copy
dioxide (CO2), 1, 165, 273, 420, 450
processing, 381 metabolic flux, 102, 103
Biomethane, 465, 466, 469, 472 rich raw materials, 274
Biomolecules, 54, 98–100, 174 corn, 276
Bionic acid, 28 sugarcane, 276
Bioprocesses, 98–100, 271, 297, 455 source, 7, 22–25, 29, 32–34, 54–57, 59,
Bioreactor, 14, 142, 216, 221, 234, 235, 65, 67, 68, 70, 73, 74, 85, 274, 393,
237, 238, 241–248, 250–255, 257, 340, 446, 455, 480
353, 354, 357, 433, 449, 486 Carbonate compounds, 11
design, 235, 253, 257 Carboxylic acids, 32
enzymatic hydrolysis process, 244 Catabolism, 29, 30, 32, 59, 74, 447
operation mode and geometric configu- Catalytic zinc-binding motif, 60, 71
ration, 244 Cell
process conditions, 247 collection, 486
fermentation process, 241, 250 cycle, 67, 353
geometric configuration and instru- densities, 3
mentation, 250 division, 11, 105, 145
process conditions, 253 equilibrium, 61
Biorefineries, 16, 36, 98, 100, 107, 123, filtration, 220
195, 196, 199, 213–215, 255, 272–277, functions, 142
279–283, 287–297, 316, 340, 344, 350, growth, 13, 14, 57, 77, 106, 110, 118,
364, 371–377, 379, 383, 386, 388, 391, 216, 218, 219, 255, 455, 486
393, 394, 401, 405, 414, 478, 480, 486, 487 harvesting, 14
Biosynthesis, 59, 67, 76, 79, 103, 107, 108, mass, 12, 143, 145, 150
111, 115, 147–149, 179, 431, 432 membrane, 28, 86, 118, 142
Biosynthetic motility, 77–79
gas, 373 population, 13
genes, 76 recycling, 3, 14, 281
Biotechnological applications, 28, 75 repair, 67
Biotechnology, 28, 99, 287, 340, 359, 364 separation, 14
evolution, 359, 364 wall, 77–79, 107, 118, 119, 146–149, 174,
Biotin, 27 181, 236, 419–426, 429, 431, 432, 434,
Bluewater footprint, 405 436, 486
Bottlenecks, 100, 102, 107, 277, 297, 463, integrity (CWI), 147
486 Cellobiose, 5, 45, 49, 83, 85, 119, 122, 178,
Brown rot fungi, 180, 411, 436 183, 235, 239, 344, 356, 389, 422, 430
Author Copy
proteins, 110 156, 178, 183, 434
recycling batch fermentation, 14 thermocellum, 43–46, 53, 54, 80–84, 87,
stress, 9 97–99, 108, 109, 119, 120, 141, 144,
thiol-redox pathways, 61 156, 178, 434
Cellulases, 46, 48, 100, 183, 238, 239, 244, genetic regulation, 83
248, 391, 425, 429, 430, 433 physiology, 46
Cellulolytic Clustered regularly interspaced short palin-
bacteria, 45 dromic repeats (CRISPR), 102, 153
enzymes, 238, 248, 344 Co-factor balancing, 98
activities, 248 Combinatorial method, 152
Cellulose Computer science, 79
hemicellulose Consolidated
degradation, 429 bioconversion process, 182, 183
hydrolysis, 183, 246, 250, 283, 351, 420 bioprocessing (CBP), 26, 43, 48, 50, 80,
Cellulosic 142, 245, 286, 297, 419, 433, 437
agricultural residues, 25 Continuous
biofuels, 475 membrane fermenter separator (CMFS), 224
ethanol, 170, 286, 287, 339–341, 345, stirred tank reactors (CSTR), 252
354–359, 362–364 Conventional
biomass pretreatment, 345 distillation, 197, 199, 201, 203, 209, 210,
production, 170, 339, 341, 345, 354, 213, 216
355, 357, 358, 362 dry-grind process, 277
Cellulosomal enzymes, 80 separation, 195, 197, 199, 201
Cellulosome, 43, 45, 48, 80, 81, 83, 85, 119, solvent, 204–206
430 Copolymers, 211
genetic regulation, 80 Corn gluten
Central metabolism, 5, 29, 46 feed (CGF), 168
Chaperones, 102, 105 meal (CGM), 168
Chemical Corrosion, 203, 205, 250, 286, 348, 382, 485
hydrolysis, 3 Cost-effective bioproducts, 98
stress, 64 Crabtree
Chemotaxis, 76 negative competitors, 35
China National Offshore Oil Corporation positive
(CNOOC), 475 species, 35
Chitosan (CS), 211, 225 yeast, 35
Chlorella vulgaris, 174, 175, 481, 486 Cross-protection, 9
Author Copy
374, 484 Enolase, 69
Dehydration, 115, 200, 201, 206, 210–214, Entner-Doudoroff (ED), 21, 74, 115, 179, 432
288, 289, 409 Environment parameters, 8
Dehydrogenase, 30, 58, 69, 102, 105, 108, Environmental
115, 143 factors, 1, 35, 75, 143, 422, 424, 431
Delayed simultaneous saccharification and pollution, 107, 141, 142, 161, 185, 234,
fermentation (DSSF), 340, 354 315, 363
Deodorizers, 14 stress response (ESR), 9
Detoxification, 57, 59, 61, 63–65, 70, 77, Enzymatic
105, 115, 123, 359 antioxidant defense system, 106
Dextrins, 278 binding, 247
D-glucopyranose molecules, 422 catalysis, 247
Dietary fibers, 345 hydrolases, 246
Dihydroxyacetone hydrolysis, 3, 5, 13, 14, 180, 181, 183,
kinase, 69 184, 233–236, 238, 240, 244, 246–248,
phosphate (DHAP), 32 255, 257, 283–286, 291, 296, 340, 346,
Diploidization, 151 348, 350, 353–357, 360, 371, 381, 385,
Direct microbial conversion (DMC), 48, 183 386, 388, 389, 391, 394, 403, 408, 412,
Disaccharides, 7, 86 413, 425, 481, 484
Distillation, 10, 45, 167, 174, 195–200, pretreated solid biomass, 389
203, 204, 206–208, 210, 212–216, 222, yield calculation, 389
277–279, 282, 288, 289, 349, 447, 486 Enzyme
Distillers dried grains (DDGs), 279, 297 cocktail, 287, 288, 388, 390, 484
Down-regulating genes, 98 expression, 65
Downregulation, 107, 149 saccharification, 386
Downstream process, 99, 216, 382 Equilibrium, 98, 197, 203, 206, 208–210,
Dry 212, 217, 246, 247
distillation grain, 289 Ergot alkaloids, 13
weight basis, 167, 174 Error-prone polymerase chain reaction, 153
Dye decolorizing peroxidase (DyP), 425, 428 Escherichia coli, 23, 25, 78, 99, 101, 102,
177, 178
E Ethanogenic bacterium, 33
Economic substrate, 122 Ethanol
Economical fluctuations, 98 adaption (EA), 144
Ehrlich pathway, 29, 30 affinity, 208
Electric flux, 47 aqueous solution, 203, 209
Author Copy
412, 413, 433 Ethanol-tolerant mutant strains, 110, 118
biochemistry, 3 Eukaryotes, 6, 153
Clostridium thermocellum, 119 Eukaryotic cells, 54
first generation bioethanol, 165 European Union (EU), 3, 164, 466, 471,
fourth generation bioethanol, 175 475, 479, 487
genetic engineering, 100 Exogenous
intermediate enzymes, 57 efflux pumps, 103
key genes, 68 enzymes, 49
organic feedstocks, 165 Exoglucanases, 5, 239, 248, 430
Saccharomyces cerevisiae, 56, 110 Exopolysaccharides (EPS), 145, 156
second generation ethanol, 170 Exothermic reaction, 10
strain improvement, 46 Exploratory fermentation, 391
third generation ethanol, 170 Exponential phase, 13, 325
Expression
transcription factor genes, 57
analysis, 70
Zymomonas mobilis, 112
dynamics, 75, 86
recover, 45, 184, 195–197, 199, 201, 208,
profile, 30, 77
209, 213, 216, 345, 433
Extractive distillation (ED), 195, 202, 204,
azeotropic distillation (AD), 199
205, 225
conventional distillation and separation
Extrinsic factors, 8
technology, 197
dehydration, 201
F
extractive distillation (ED), 202
nonconventional technologies, 202 Fatty acid, 1, 2, 7, 9, 11, 12, 64, 119, 144,
organophilic membranes, 208 146, 153, 436, 450, 466
selective membranes, 208 metabolism, 67, 147
separation membrane technology, 207 Feedstock, 15, 44, 45, 48, 49, 98, 99,
pervaporation (PV), 207 122, 162, 167, 170, 172, 173, 175, 176,
stress, 32, 35, 64, 73, 78, 111, 142, 143, 184, 185, 233, 235, 273, 282, 283, 286,
145, 146, 149, 154 289–291, 295, 296, 344, 382, 432, 446,
tolerance, 23, 26, 32, 35, 47, 59, 65, 78, 465, 476, 477, 479, 480, 483, 485, 487
80, 83, 85, 86, 111, 114–116, 118, Fermentation
141–147, 149–156 approaches, 339, 340, 354, 359, 360, 363,
AdhE mutation, 155 364
evolutionary engineering, 150 conversion yield, 392
genes involved, 147 microorganism used, 177
genetic approaches, 151 pathway, 34, 433
Fermentative Gene
growth, 7 activation, 62, 106
yeasts, 10 coding, 68
Ferredoxin oxidoreductase, 43, 47 encoding, 32, 46–48, 59, 61, 63, 83, 103,
Apple Academic Press
Author Copy
biofuels, 108, 170, 373, 375, 478 240
ethanol, 44, 56, 165, 287 Genetic
Flavoprotein, 76 approaches
Flex-fuel engines, 15, 199 genome editing and shuffling, 152
Fluorescent in situ hybridization (FISH), 252 global transcription machinery
Fluxomics, 65 (gTME), 153
Food industry, 2, 276 over-expression, 153
Forkhead transcription factor, 72 random mutagenesis, 151
Formic acid, 75, 374 engineering, 3, 23, 32, 49, 55, 99, 100,
Fossil fuels, 98, 122, 123, 141, 142, 161, 102, 104, 105, 107, 108, 110, 111, 122,
170, 216, 233, 234, 272, 315, 331, 371, 144, 151, 167, 279, 433, 434
372, 377, 394, 412, 419, 420, 445, 446, expression, 87
463, 464, 467, 473, 476, 477, 480, 487 mechanisms, 59
Fractionation, 172, 174, 375, 436 mutation, 9
Freeze-thaw treatment, 151 perturbations, 152
Fructans, 73 regulation, 53–56, 59, 67, 73–75, 77, 78,
Fructose, 7, 27, 34, 69, 325, 343, 432, 433 80, 81, 83, 85, 87
Fumaric acid, 65 stability, 181, 435, 486
Functional Genome, 3, 6, 31, 54, 59, 74, 75, 86, 106,
group, 144, 206 113, 115, 117, 118, 145, 150, 152–155
protein association networks, 69 analysis, 83
Fungal editing, 106, 152, 154
biomass, 13 Genomics, 65, 85, 146
cells, 13 Geothermal areas, 86
plasma membranes, 10 Global
Fungi, 5, 10, 11, 13, 99, 108, 141, 180, 284, bioenergy partnership (GBEP), 475
340, 353, 411, 421, 422, 425–432, 434–436 metabolic networks, 53
Furan aldehyde stress, 65 sorghum production, 319
sustainable goals, 487
G transcription machinery, 111, 124, 153
Galactans, 423 engineering (gTME), 111, 124
Galactomannans, 423 warming (GW), 142, 234, 401, 467, 487
Gasoline, 2, 8, 15, 162–164, 196, 199, 201, Glucan, 348, 349, 351, 352, 355, 375,
206, 234, 272, 316, 325, 409, 412, 420, 388–392, 423, 481
464, 466, 467, 474, 475, 484 Glucokinase, 58
Author Copy
Glycerol
Hexokinase, 58, 69, 143
3-phosphate (G3P), 32
Hexoses, 5, 7, 23, 24, 86, 107, 165, 167,
assimilation, 24
176, 181, 183, 184, 236, 240, 250, 253,
pyruvic fermentation, 24
254, 286, 340, 374, 403, 432, 433, 435,
Glycine, 64, 147, 425
Glycogen, 9, 11 485
Glycolaldehyde, 63 High
Glycolysis, 5, 24, 29, 30, 46, 57, 64, 67, 68, osmolarity glycerol, 32
73, 75, 85, 110, 122, 432 performance liquid chromatography
Glycolytic genes transcriptional activator, 72 (HPLC), 379
Glycosidases, 485 Homeostasis, 6, 11, 61, 67, 69, 143, 146, 153
Glycoside hydrolases, 430, 485 Homoethanol pathway, 75
Glycosylation, 55 Homogenization, 247
G-protein, 149 Homogenous azeotropic distillation (HAD),
Gram-negative 199
anaerobic bacterium, 74 Homologous
bacteria, 27, 112, 143 expression, 485
Gram-positive recombination (HR), 67
anaerobic bacterium, 119 Hopane, 28
bacterium, 143 Hoponaid (hpn), 151
thermophilic microbe, 45 Hybrid
Greenhouse gas (GHG), 162, 234, 331, 409, distillation-pervaporation (PV) columns,
464, 487 213
Growth separation, 213
conditions, 12, 16, 26, 83, 154 Hydrocarbon, 201, 372, 420
inhibitor, 10, 75 Hydrochloric acid (HCl), 284, 297
Hydrodynamic cavitation (HC), 242–244,
H 284, 482
Heat reactors (HCR), 243
shock, 10, 35, 149 Hydrogen
factor protein, 62, 72 evolution reaction (HER), 452, 455
proteins (HSP), 10, 35, 36, 75, 102, peroxide (H2O2), 238, 244, 425, 437, 482
149 Hydrogenase deletion, 47
stress, 102 Hydrolytic
transcription factor, 32, 35 enzyme, 44, 45, 48, 119, 352, 388
stress, 9, 72, 75, 105, 106 linkage, 44
Hydrothermal
liquefaction reactors, 241 metabolites, 118
pretreatment, 351, 371, 381–385, 391 redox potential, 105
fundamentals, 381 Ionic
processing, 381, 394 liquids (ILs), 204, 205, 222, 284, 350, 409,
Hydroxycinnamoyl transferase (HCT), 107 414
Hydroxymethylfurfural, 238, 348, 374, 403 stress, 61
Hyperbranched polymer (HyPol), 196, 206, Iron-sulfur clusters (ISC), 153
207, 225 Isoamyl acetate, 12
Author Copy
Isobutanol, 30, 45, 47, 102, 104, 105
I Isoleucine, 29, 64, 103, 147
Isomerase pathway, 112
Immobilization, 219, 221
In situ J
ethanol recovery
hybrid fermentation technology, 216 Jalview software, 60, 71
product recovery (ISPR), 216
Incubation, 27, 99, 151, 180, 325
K
period, 180 Keto acids, 29
Industrial Ketones, 6, 61, 63, 219
fermentation, 2, 12, 36, 105, 142, 433 Klebsiella pneumoniae, 103
scientific research council (CSIR), 378 Kluyveromyces marxianus, 3, 4, 54, 65–73,
Industrialization, 404, 419 87, 112, 256
Inhibitor, 3, 12, 14, 57, 59, 61, 63, 64, 73, genetic regulation, 67
75, 77, 78, 104, 115, 118, 123, 141, 181, metabolomic analysis, 73
238, 240, 283, 284, 295, 345, 352, 382, molecular strategies, 67
403, 408, 481
compounds, 110, 484 L
tolerance, 104, 108, 181 Laccases, 388, 425, 429, 432
genetic engineering, 104 Lactate dehydrogenase (LDH), 43, 46, 47,
Inhibitory compound, 57, 59, 65, 73, 246, 50, 121
295, 348 Lactic acid, 10, 43, 374
stress, 77 Lag phase, 13, 104
Inner membrane proteins, 103 Lallemand biofuels, 56
Inorganic membranes, 209–212 Laminaribiose, 81, 83
Integrated Large-scale fermenters, 16
corn, 288 Leucine, 29, 103
high gravity (iHG), 409 Life cycle assessment (LCA), 467, 487
Intelligent microbial heat-regulating engine Lignin, 107, 174, 235, 236, 242, 387, 403,
(IMHeRE), 106 421, 425, 426
Intensifiers, 70 degrading peroxidases, 426
Interfacial tension, 218 dye decolorizing peroxidase (DyP), 428
Intermediary lignin peroxidase (LiP), 426
enzymes, 68 manganese peroxidase (MnP), 427
molecules, 55, 64 versatile peroxidase (VP), 428
Author Copy
Matrix, 178, 211, 283, 373–375, 402, 408,
339, 374, 401, 405, 406, 413, 414, 478 474, 475, 480, 481
overview, 374 Mechanical resistance, 107
ethanol production, 53, 70, 240, 243, 403, Membrane
404, 412, 414 assisted vapor stripping (MAVS), 216, 225
feedstocks, 142, 295, 345, 408 biogenesis, 77–79
hydrolysates, 12, 53, 64, 65, 115 composition, 119, 145, 147
materials, 3, 5, 7, 8, 13, 44–46, 49, 104, organization, 147
108, 119, 170, 175, 176, 178, 180, 235, permeability, 111, 145, 146
236, 238, 246, 248, 274, 283, 284, 286, permeabilization, 145
291, 295, 340, 345, 346, 348–352, 354, perturbation, 115
374, 375, 381, 394, 407, 408, 411, 419, selectivity, 207, 208, 210–212, 225
421–425, 432, 433 transporters, 103, 112
cellulose, 422 Metabolic
extractives and ashes, 424 activities, 9, 102, 110, 142
hemicellulose, 423 analysis, 64, 80
lignin, 421 branching, 6
pectin, 424 engineered microorganisms, 110, 123
pretreatments, 107 engineering, 29, 36, 46, 75, 80, 85,
residues, 25, 36, 63, 283, 376, 419 97–101, 113, 115, 116, 118–120, 123,
substrates, 16, 83, 253, 436 124, 172, 433
Lipid, 46, 57, 73, 144, 147, 151, 174, 374, fluxes, 68
480, 481 function, 70, 102
peroxidation, 106 heat, 106
Liquefaction, 181, 242, 243, 277, 278, 355, pathway, 6, 13, 24, 28–30, 32, 33, 54, 55,
359, 403, 472 57, 64, 67, 74, 77, 86, 98, 119, 122,
Liquefied biogas (LBG), 472 123, 143, 167, 177, 455
Liquid reactions, 80, 122, 219
hot water (LHW), 180, 351, 382, 383, regulation analysis, 68
413, 485 Metabolism, 5, 7, 11, 13, 21, 28–31, 35, 57,
liquid extraction (LLX), 197, 217–222, 225 59, 63, 67, 68, 73, 76, 78, 85, 99, 104,
phase, 11, 199, 203, 385 110, 115, 119, 123, 143, 378, 485
hydrolysis, 5 Metabolome, 64, 73
Liquor tax imposition, 162 Metabolomic, 64, 65, 85, 99, 122, 144, 146
Logarithmic cells, 13 analysis, 64, 73, 146
Long-term productivity, 14 data, 79
Microalgae, 174, 177, 184, 273, 373, 465, 486 tools, 100, 106, 142
Microalgal weight, 64, 423
biomass, 174, 181 Monoethylene glycol, 282
cells, 174, 181 Monomers, 44, 239, 379, 380, 382, 384,
Microarray, 110 385, 388, 402, 422, 429, 430, 432
analysis, 78, 111 Monosaccharides, 100, 165, 174
Microbes, 45, 98, 141, 142, 147, 176, 295, Morphological abnormalities, 105
434, 451 MspI gene, 106
Microbial Mucor indicus, 25, 356
Author Copy
bioeletroshyntesis (MES), 447 Multidrug resistance transporters, 63
cells, 53, 144 Multienzymatic systems, 80
consortium, 122, 176, 237, 358 Multiple
desalination cell (MDC), 448, 449, 455 genetic tools, 36
electrodialysis cell (MEDC), 448, 455 metabolic pathways, 55, 64
electrolysis cell (MEC), 447–450, 455 sequence alignment analysis, 59, 60, 70, 71
electrosynthesis (MES), 445, 448, 449, Multi-round atmospheric and room tempera-
451, 455 ture plasma (mARTP), 115
fermentation, 141, 142, 162, 174, 253 MUSCLE algorithm, 60, 71
fuel cell (MFC), 447–449, 455 Mutagenesis, 4, 9, 49, 75, 104, 105,
growth, 23, 26, 35, 104, 295, 388, 452 115–119, 144, 146, 151–153
heat shock stress, 35 Mutants, 101, 115, 118, 148, 149
reverse electrodialysis electrolysis cell Mutation, 111, 115, 118, 151, 152, 155, 485
(MREC), 448 Mutational pathways, 150
strain, 7, 105, 141
tolerance, 47, 104 N
transcriptional profiles, 102, 153
Microbiological control agent, 10 Nanoparticle, 212
Micronutrients, 6, 27 National
Microwave, 241, 283, 284, 339, 340, 352, Biodiesel Production and Use Program
353, 358, 360, 408 (PNPB), 474
irradiation reactor (MWIR), 241, 242 Research Council (NRC), 405
reactors (MWR), 241 Native xylokinase, 68
Milling, 241, 277, 283, 294, 371, 408, 481 N-butanol tolerance, 103, 155
Mitochondria, 6, 57, 104 Near-universal stress factor, 142
Mitochondrial Negative regulation, 54, 55
enzymes, 68 Next-generation sequencing (NGS), 75
isozyme, 57 Nicotinamide
transcription factor, 72 adenine dinucleotide, 76
Molasses, 164, 167, 198, 273, 274, salvage pathway, 105
279–281, 289, 317 Nitrogen
Molecular metabolites (NCR), 30
biology, 31, 99 source, 1, 325
mechanism, 61, 64, 147, 150 Nitrogenous compounds, 8
methods, 3 Nonconventional solvents, 202
Author Copy
sequence site, 74 176, 181, 183, 184, 236, 250, 254, 286,
Nucleus, 55 340, 374, 403, 432, 433, 485
Nutrient starvation, 72 phosphate pathway (PPP), 67, 104
Periplasmic linkers, 103
O Permeability, 24, 62, 110, 111, 144, 146,
207, 208, 219, 225
Octadecanoic acids, 64
Peroxiredoxin, 69
Oligosaccharides, 181, 278
Pervaporation (PV), 196, 207, 208,
Omics, 85, 87, 99, 146, 151, 154, 156
212–215, 223
technique, 73, 86, 87, 156
Pervaporative fermentation, 197, 222
Optimal growth temperature, 10, 105
Pesticides, 14, 317, 328, 483
Organic
pH, 4, 8, 10, 11, 13, 24, 27, 31, 33, 34, 45,
acids, 10, 57, 61, 65, 67, 100, 120, 180,
61, 63, 74, 86, 110, 111, 115, 118, 121,
218, 222, 281, 350, 481
143, 146, 147, 153, 181, 182, 209, 218,
compound, 5, 447, 451
interface, 218 225, 234, 239, 240, 243, 244, 246–249,
membranes, 210 251, 253, 255, 277, 278, 322, 325, 326,
nitrogen, 8 329, 349, 355, 356, 384, 387–389, 392,
solvents, 102, 218, 219, 339, 340, 349, 350 409–411, 426–429, 431, 451, 452, 480
Organization of Arab petroleum exporting Pharmaceutical, 99
countries (OPEC), 162 products, 14
Organophilic membranes, 208, 222 Phenolic
Osmoprotectants, 147 alcohols, 77
Osmostress response, 62 aldehyde, 75, 77, 78, 118
Osmotic adaptation, 78
pressure, 32, 35, 144, 423 stress, 78
stress, 23, 24, 34, 72, 240 compounds, 1, 2, 104, 105, 427, 429, 435
Osmotolerance, 100 inhibitors, 63
Overexpression, 30, 57, 67, 68, 76, 85, 86, Phenols, 57, 61, 65, 75
100, 102–105, 108, 111–114, 116, 118, Phenotype, 8, 64, 111, 142–144, 146, 151,
119, 149, 436 153–155
Oxidation, 5, 6, 57, 63, 68, 70, 180, 409, Phenylalanine, 29, 73, 236, 422, 432
426–429, 432, 445, 449, 451 Phlegma streams, 198, 213, 214
Oxidative stress, 35, 58, 61, 63, 72, 75, 104, Phosphatidic acid (PA), 155
105, 144, 149, 236, 428 Phosphatidylcholine, 73, 146, 155
Oxidoreductase, 6, 104, 105, 112, 121, 429 Phosphatidylethanolamine, 58, 65, 73
Oxygen surface tension, 11 Phosphoenolpyruvate carboxykinase, 69, 120
Author Copy
Pigments, 1, 2, 99, 425 Pyruvate
Plant cell wall biosynthesis, 431 carboxylase, 100, 121
Plasma membrane, 11, 61, 73, 111, 144, decarboxylase (PDC), 6, 68, 69, 115, 121
146, 219 ferredoxin oxidoreductase, 46, 47, 85
Plasmids, 31, 75, 103 kinase, 69, 105, 117, 120
Pleiotropy, 153 Pyruvic acid, 24, 28
Polar
flagella, 27 Q
solvents, 217 Quorum-regulating system, 106
Poly(1-trimethylsily-1-propyne) (PTMSP),
209 R
Poly(ethylene glycol), 206
Polydimethylsiloxane (PDMS), 209, 225 Raw material, 2, 12, 14, 16, 22, 32, 53, 59,
Polyglycerol, 206, 207 77, 108, 141, 161, 168, 184, 234, 240,
Polymer, 29, 99, 107, 206, 209, 235, 236, 273–277, 287, 289, 292–294, 315–317,
238, 277, 284, 374, 375, 388, 402, 407, 321, 323, 325, 339, 345, 371, 373–375,
421–425, 431, 436, 452 377–379, 381, 382, 391, 407, 408, 411,
Polymeric membranes, 209, 210, 213 412, 425, 454, 465, 474, 476, 484
Polypeptide, 55, 76 Reactive oxygen species (ROS), 57, 113,
Polysaccharide, 9, 32, 81, 83, 178, 236, 248, 144, 482
382, 423, 424, 430, 431 Recalcitrance, 107, 344, 345, 375, 377, 421,
Portable digital refractometer, 329 425, 432, 478
Positive regulation, 54 Redox balance, 6
Posttranslational regulation, 68 Reflux ratios, 196, 197
Potassium aluminosilicates, 196, 201 Renewable
Potential protein-encoding genes, 31 diesel blendstock (RDB), 408
Prefoldin (PFD), 108, 124 energy, 22, 26, 29, 46, 316, 373, 377, 402,
Pre-saccharification, 26, 340, 355, 371, 392 446, 455, 463, 465, 467, 469, 472, 473,
pre-saccharification and fermentation 475
strategy (PSSF), 392 sources, 44
Pressure swing adsorption (PSA), 201 fuel
Pretreatment methodologies, 347, 364 association, 3, 375
Primary yeast taxa, 56 program (RFS), 475
Principal ethanol-producing yeasts, 2 Resistant yeast, 63
Production systems, 324, 331 Respiratory growth, 7
Prokaryotes, 28, 153 Restriction-modification system, 106
Roger, 326
Rot fungi, 180, 284, 411, 426, 427, 434–436 aliphatic acids, 104
Signaling pathways, 67, 68, 146
S Simultaneous saccharification and,
co-fermentation (SSCF), 13, 182, 183,
Saccharides, 7, 278
245, 297, 340, 354, 358, 433
Saccharification, 9, 13, 26, 59, 105, 108,
fermentation (SSF), 13, 26, 105, 182,
174, 182, 183, 244–248, 250, 277, 278,
183, 245, 278, 340, 354, 433
284, 295, 297, 340, 345, 347, 354, 355,
Small RNA (sRNA), 32, 103, 154
Author Copy
357–359, 361, 363, 364, 371, 375, 382,
Soaking aqueous ammonia (SAA), 180
386, 388, 389, 392, 409, 433, 434, 481
Socio-economic troubles, 22
and fermentation, 9, 13, 26, 105, 182,
Soft rot fungi, 180, 436
183, 245, 246, 250, 278, 297, 339, 340,
Solid
354, 355, 357–359, 364, 392, 433, 434
biomass, 389
Saccharomyces, 2–4, 7, 21, 24, 29–31, liquid interfaces, 5
33–36, 44, 54, 56, 59–62, 64, 87, 98, 99, Solvent
101, 103, 108–111, 119, 141, 143, 144, stress, 103
167, 177–179, 181, 240, 252, 253, 255, tolerance, 102
278, 281, 286, 317, 325, 328, 355, 391, Sorghum, 66, 166, 167, 249, 254, 256, 273,
392, 432 276, 315–327, 331, 339–364, 375
cerevisiae, 2–4, 7, 21, 29–31, 33–36, 44, bioenergy production, 344
54, 56, 59–62, 64, 87, 98, 99, 101, 103, sweet sorghum bagasse, 343, 345
108–111, 119, 141, 143, 167, 177–179, world production, 341
181, 240, 252, 253, 255, 278, 281, 286, Spt-Ada-Gcn-5 acetyltransferase, 154
317, 325, 328, 355, 391, 392, 432 Stagnant fermentations, 12
adaptation and tolerance, 61 Stain fungi, 436
genetic regulation, 59 Starch integration, 294
metabolomics analysis, 64 Stationary phase, 13, 70, 325, 329
stress tolerance, 64 Steam explosion (SE), 180, 237, 284, 351,
uvarum, 7 382, 410, 413, 485
Saturation, 252, 372 Sterols, 7, 9, 11, 12, 144
Secondary Stirred tank
fungal metabolic compounds, 13 bioreactors (STBR), 243
metabolism, 13 reactor, 252
Second-generation Stoichiometric conversion, 241, 254
bioethanol, 16, 22, 108, 170, 250, 283, Strain improvement, 43, 46, 50, 151, 152
284, 286, 288, 297, 321, 373, 376, 391, Stress
393, 394, 446, 476, 484 conditions, 10, 34, 58, 59, 67–69, 100,
biorefineries, 386, 388 110, 111, 149
development, 375 factors, 55, 57, 142, 150, 154
Selective milling technology (SMT), 287 response program, 9
Separate hydrolysis and, tolerance, 64, 78, 146, 153, 343
co-fermentation (SHCF), 26 Substrate utilization, 6
fermentation (SHF), 13, 26, 182, 245, Succinate dehydrogenase, 69, 103
278, 297, 354, 356, 421, 437 Succinic acid, 45
Author Copy
Surface tension, 8, 11 bacteria, 86, 145
Surplus, 7, 296 cellulolytic
Suspended solid particles, 372 bacteria, 49
Sustainable microorganisms, 45
alternative fuels, 2 Thermo-protection, 10
biofuels, 100 Thermo-robustness, 106
genetic engineering, 100 Thermostability, 105, 106, 123
economic progress, 372 genetic engineering, 105
energy, 474 Thermotolerance, 65, 105, 106
power source, 98 Thermotolerant, 10, 35, 73, 105, 106, 111,
renewable natural sources, 25 254, 355
scenario, 473 Thiamine metabolism regulatory protein, 58
Sweet sorghum, 254, 315–327, 331, Thiol-specific peroxidases, 61
339–341, 343–364 Thioredoxin, 61, 68
advantages, 320 Threshold maltose, 7
bagasse, 249, 339, 340, 345–359, Titanium dioxide (TiO2), 212
362–364 Tolerance, 1–3, 7, 12, 23, 24, 26, 27, 31–35,
biology, 318 47, 59, 61, 63–65, 74, 78, 80, 83, 85, 86,
basic races, 318 102–106, 108, 110, 111, 113–118, 123,
intermediate, 319 141–147, 149–156, 176, 181, 216, 240,
conversion, 325 316, 321, 340, 343, 486
production, 321, 324 mechanism, 80, 146, 155
energy use, 324 Toxic
varieties, 315, 317, 320, 326, 331, 343 agents, 77
Synthetic biology, 29, 59, 98, 100, 106, 123, compounds, 57, 58, 65, 74, 76, 103, 117,
124, 156 118, 347, 435
Syringaldehyde, 77, 374 metabolites, 13
molecules, 67
T Traditional distillation schemes, 196
TALEN assisted multiplex editing (TAME), Transaldolase (Tal), 30, 68
152 Transamination reaction, 29
Target mRNA interactions, 32 Transcription
Technological challenges, 272, 294, 463, activator-like effector nuclease, 152
478, 487 factor (TFS), 57–59, 61, 62, 70, 72, 111,
Temperature tolerance, 35 113, 116, 149, 153, 154
Transcriptional W
activation, 149
Water
activator protein, 62, 72
consumption
factors, 64
biological pre-treatment, 411
regulation, 76–79
Apple Academic Press
Author Copy
475, 480, 482, 487 Wild yeasts, 2
Trehalose, 9, 11, 32, 36, 65, 105, 111, 143,
147, 149, 151 X
6-phosphate
phosphatase (TPP), 32, 36 Xylan, 49, 84, 236, 348, 349, 351, 352, 374,
synthase (TPS), 32 375, 388, 423, 431
degradation, 111 Xylitol dehydrogenase (XDH), 68, 112
Tricarboxylic acid cycle, 100 Xylogalacturonan (XGA), 424
Tryptophan, 73, 111 Xyloglucan, 423
Xylooligosaccharides (XOs), 374
U Xylose, 3, 24, 34, 46, 49, 68, 73, 105, 107,
110, 112–115, 119, 122, 165, 167, 178,
Ubiquitin, 105 181, 236, 240–242, 246, 250, 286, 344,
Ultrasound-assisted treatment, 352 348, 358, 374, 380, 423, 424, 432, 433
United adaptation, 68
Nations (UN), 319, 341, 372 isomerase (XI), 112, 122
States (US), 3, 15, 22, 23, 164, 165, 209, reductase (XR), 68, 112
273, 275, 287, 342, 406, 446, 471, 473, heterologous (XYL1), 68
479 Xylulokinase, 112, 122
Uronic acids, 107, 374, 403
Y
V Yeast
Vacuolar acidification, 145, 146 cells, 2, 10, 11, 14, 16, 24, 63, 111, 143,
Vacuole membranes, 104 147, 150, 197, 219–224, 282
Valine, 29, 64, 65, 103, 114, 147 physiology, 8, 12, 16, 57
Vanillin, 77, 117, 181
Vapor-phase adsorption, 201 Z
Variety ROGER, 315, 326, 328, 329, 331 Zea mays, 166, 276
Vegetable biomass, 106 Zeolite, 201, 209, 210, 212, 216, 219
molecular tools, 106 adsorption units, 201
Velocity, 247 Zero population growth, 13
Versatile peroxidase (VP), 426, 428 Zinc
Viscosity, 206, 217, 218, 244, 247, 356, 485 finger
Volatility, 2, 196, 197, 202, 203, 206, 207, 411 nuclease (ZFN), 152
Volumetric productivity, 12 protein, 72
Vortexes, 246 starvation, 58
mobilis, 21, 26–31, 33, 34, 44, 48, 53, 54, 77, 78
74–79, 87, 97, 99, 102, 108, 109, 112, overview, 26
115, 121, 141, 144, 156, 177, 178, 181, physiological advantages, 30
254, 358
Author Copy