You are on page 1of 52

Journal Pre-proof

An overview of integration opportunities for sustainable bioethanol production from


first- and second-generation sugar-based feedstocks

Bamidele Victor Ayodele, May Ali Alsaffar, Siti Indati Mustapa

PII: S0959-6526(19)33727-8
DOI: https://doi.org/10.1016/j.jclepro.2019.118857
Reference: JCLP 118857

To appear in: Journal of Cleaner Production

Received Date: 01 November 2018


Accepted Date: 11 October 2019

Please cite this article as: Bamidele Victor Ayodele, May Ali Alsaffar, Siti Indati Mustapa, An
overview of integration opportunities for sustainable bioethanol production from first- and second-
generation sugar-based feedstocks, Journal of Cleaner Production (2019), https://doi.org/10.1016/j.
jclepro.2019.118857

This is a PDF file of an article that has undergone enhancements after acceptance, such as the
addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive
version of record. This version will undergo additional copyediting, typesetting and review before it
is published in its final form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier.


Journal Pre-proof

An overview of integration opportunities for sustainable bioethanol


production from first- and second-generation sugar-based feedstocks

Bamidele Victor Ayodelea*, May Ali Alsaffarb, Siti Indati Mustapaa


a Institute of Energy Policy and Research, University Tenaga Nasional

Jalan IKRAM-UNITEN 43000, Kajang Selangor, Malaysia


bDepartment of Chemical Engineering, University of Technology Iraq

Corresponding author email: ayodelebv@gmail.com, ayodele.victor@uniten.edu.my


Journal Pre-proof

Sugarcane Milling Juice


Steam

Bagasse Juice treatment


CHP Electricity
To Grid
Steam explosion Pre- Clarification
treatment Lignin
CO2 Continuous
Pre-treated liquor feeding

Pre-treated slurry SSF


(Pressed or whole slurry) Fed-batch Distillation

Xylose Enzyme
fermenting
Bioethanol
Yeast
Journal Pre-proof

An overview of integration opportunities for sustainable bioethanol


production from first- and second-generation sugar-based feedstocks

Nomenclature

1G First generation

2G Second generation

1G2G Integrated first and second generation

WIS Water insoluble solid

CHP Combined heat and power

CBP Consolidating bioprocessing

SSF Simultaneous saccharification and fermentation

SHF Separate hydrolysis and fermentation

SSCF Simultaneous saccharification and co-fermentation

EH Enzymatic hydrolysis

EHE Enzymatic hydrolysis efficiency

MESP Minimum ethanol selling price

IRR Internal rate of return

SPWS Steam pre-treated wheat straw

PWM Pre-saccharified wheat meal

Abstract

The production of bioethanol from second generation feedstocks which are mainly
lignocellulosic biomass provides the opportunities for a cleaner, and low carbon biofuel that
can serve as an alternative to fossil fuel. However, one major constrain is the high cost of
production due to expensive pre-treatment technologies. On the other hand, bioethanol
production using first generation feedstock is a proven and well-established technology with
high bioethanol productivity and yield, but the process is identified with the issues of food-to-
fuel debate and high environmental impact from land use charge. Integrating second generation
bioethanol production process with the well proven first generation bioethanol production
facilities has the potential of harnessing the synergistic effect that could maximize the technical,

1
Journal Pre-proof
economic and environmental benefits of the integrated first and second generation process. This
study examined literature on studies related to first generation bioethanol production, second
generation bioethanol production and the integrated first and second generation processes with
the aim of identifying the extent of integration opportunities for first and second generation
bioethanol production from different sugar-based feedstocks. The review put into consideration
the different process scenarios available in terms of the technical, economic and environmental
benefits of the bioethanol production processes. The benefits of the different integration
scenarios for bioethanol production were considered in comparison with the standalone first
generation and second generation bioethanol production scenarios. Based on the consideration
of the different integration scenarios, some recommendations were proposed for attention in
future research on the integration of first and second generation bioethanol production from
sugar-based feedstocks.

Keywords: Bioethanol; First generation; Integration opportunity; Second generation;


Sugarcane

1. Introduction
Cleaner energy production from biomass materials has been identified as one of the ways
through which overdependence on fossil fuels can be reduced (Lozano and Lozano, 2018).
According to the United State Energy Information Administration (EIA) 2017 report, petroleum
and other liquid fuels are the dominants sources of transportation energy globally (EIA, 2017).
The utilization of these fossil-based fuel as sources of transportation energy have to large extent
contributed to emission of carbon dioxide (CO2) and other greenhouse gases (Fan et al., 2018).
Biofuel production from biomass resources is one of the preferred routes through which the
transportation sector can be made “green” (Kumar et al., 2019). As a result of this, the demand
for biofuel in the form of bioethanol, biobutanol and biodiesel has been on the increase and thus
there is need for a corresponding increase in its production in order to meet this ever increasing
demand (Uría-martínez et al., 2018). Amongst the aforementioned biofuels, bioethanol is the
biofuel with the largest global production capacity and is also the most advanced in terms of
government policies on biofuel blend with gasoline (Niphadkar et al., 2018). This has made
bioethanol the most demanded biofuel at the moment (Renewable Fuels Association, 2019).

Being the most demanded biofuel; bioethanol production has passed through several
technological advancements which has increased global production capacity (Gavahian et al.,
2018). Most of the commercial bioethanol production plants depend on sugar-and starch-based
feedstocks, such as corn in the USA (Swain and Mohanty, 2019), sugarcane in Brazil (Paulino
et al., 2018) and wheat, sugar beet and barley in Europe (Friedl, 2019; Marzo et al., 2019).
These modes of bioethanol production from sugar- and starch-based raw materials are also
known as first generation (1G) production (Damay et al., 2018). 1G bioethanol production from
sugar-based feedstocks promises to be a good replacement for petroleum-derived transportation
fuel due to its lower pollutant emission compare to corn-based bioethanol (Dutta et al., 2014).
Countries such as Brazil and US have been at the fore front of bioethanol production for more
than two decades (Mekonnen et al., 2018). For instance the technology for the commercial
production of bioethanol from sugarcane juice has been in used for the past two decades in
Brazil (Macrelli et al., 2012). 1G bioethanol production has several advantages such as low
production cost, familiar feed stock, and energy efficient production methods, resulting in lower

2
Journal Pre-proof
fossil fuel requirements in the biofuel value chain-with resulting benefits in mitigation of CHG
emissions (Naik et al., 2010). Nevertheless, issue of 1G bioethanol production competing with
food has raised a serious debate on its sustainability (Rudel, 2013), especially because of the
demand for highly fertile soils with high rainfall and/or irrigation for sugarcane cultivation
(Finkbeiner, 2014).

More prospects on bioethanol production can be tapped using non-food lignocellulosic biomass
where large amounts of these bagasse and straw are co-produced during processing of
sugarcane juice for either bioethanol or sugar production (Hemansi et al., 2019). In Brazil
bagasse is generally used as either fuel in boilers to meet the steam energy requirement of the
1G ethanol production process or to generate electricity which are sold to the grid (Dias et al.,
2013a). Neither of these is in line with the requirement of the modern biorefinery concepts
which demands for the efficient use of the limited available biomass resources (Zhou et al.,
2018). The sugarcane bagasse and straw are lignocellulose materials which can be utilised
efficiently for the simultaneous production of second generation (2G) bioethanol, bioelectricity
and heat (Chandel et al., 2018). Efforts have been made through extensive research for the
development of viable technology for production of bioethanol from lignocellulosic biomass
(Arora et al., 2019). Despites the efforts made so far, the economic feasibility of large scale
production of 2G in stand-alone facility is still being questioned (Slupska and Bushong, 2019).
However, commercial 2G bioethanol production from corn stover has started in Italy and US
while demonstration plants for commercial scale production are under development in Sweden
and Canada (Europian Biofuel Technology Platform, 2013; Pescarolo, 2013; Slupska and
Bushong, 2019).

One of the major constraints with the 2G bioethanol production is the pre-treatment stage which
constitutes 18 % of the production cost (Rajendran et al., 2018). Other parameters influencing
2G bioethanol production are the capital cost of the plant, cost of feedstock, enzyme and energy
(Erdei et al., 2013). One of the ways proposed to reduce 2G production cost is through high
ethanol yield and concentration (Naresh Kumar et al., 2019). The later can be achieved by
increasing the amount of water-insoluble solids (WIS) (Macrelli et al., 2012). This might likely
leads to decrease in yield as a result of inhibition caused by degradation products and reduced
mass transfer (Quintero et al., 2013). Apart from pre-treatment, the other major concern is the
energy consumption of distillation (Oliveira et al., 2016). It is often difficult to attain the
minimum 40 g/L or more of bioethanol concentration in standalone 2G process, while
bioethanol concentration as high as 80 to 115 g/L can be obtained from fermentation broth of
1G process (Della-Bianca et al., 2013). The biological limitations in the 2G process results in
a more dilute ethanol product, which thus increases distillation costs compared to 1G process.
Mixing 1G and 2G sugar streams and/or fermentation products increases ethanol concentration
in feed to distillation, thus reducing energy consumption (Macrelli et al., 2012). Therefore, there
is significance scope to optimize the 1G2G fermentation process through combinations of
different sugar streams which is the main focus of this present review. Bearing all these in mind,
it will be of great significant to have an overview of the whole process (standalone 1G, 2G and
integrated 1G2G) considering bioethanol and CHP production, as well as the net available
energy. There are scanty review articles which examines the production of 1G2G bioethanol in
a single facility from different approach. Naik et al. (2010) focused on the effective technologies
and processes for the conversion of biomass into biofuels and bio-products from different
3
Journal Pre-proof
feedstocks. Damartzis and Zabaniotou (2011) presented an analysis of the thermochemical
conversion of biomass to second generation liquid biofuels and challenges of process
integrations for innovative and sustainable solutions to climate concerns and energy security.
Saladini et al. (2016) examined the application of third generation feedstocks for biofuel
production. Rastogi and Shrivastava (2017) presented a review of different pre-treatment,
sacharifiction, and fermentation processes employed for 2G bioethanol production. Bechara et
al. (2018) presented an overview of different design strategies for 1G and 2G bioethanol
production. Ferreira et al. (2018) analysed different integration strategies of combining 2G with
waste streams from 1G stream. The present review focused on the integration opportunities for
sustainable and cleaner bioethanol production from 1G and 2G sugar-based feedstocks. In the review,
an overview of the 1G bioethanol production, 2G bioethanol production, integration strategies for 1G2G
bioethanol production, techno-economics studies of integrated 1G2G and their environmental benefits
were presented. The review was concluded with some recommendation for further studies.

2. 1G bioethanol production
First generation bioethanol is generally produced using starch- or sugar-based crops such as
barley (Gibreel et al., 2009), corn (Del Carmen Fong Lopez et al., 2019) wheat (Patni et al.,
2013), sugarcane (Mączyńska et al., 2019), sugar beet (Marzo et al., 2019), and sweet sorghum
(Appiah-Nkansah et al., 2018) by means of a simple fermentation of sugar extracted from such
crops. The process for 1G bioethanol production varies considerably depending whether it is
starch or sugar-based feedstock (Mohanty and Swain, 2019). However, most of the main
process stages remain the same despites their differences in the process conditions such as
temperature and pressure. The process steps for starch-based bioethanol production include
milling, gelatinization, hydrolysis, fermentation and distillation (Szambelan et al., 2018). While
bioethanol production from sugar-based feedstocks include all the process steps in starch-based
except for gelatinization and hydrolysis (Gomez-Flores et al., 2018). Gelatinization and
hydrolysis are included in starch-based bioethanol production process because the carbohydrate
chains in starch need to be hydrolysed to glucose for fermentative conversion to bioethanol
(Bothast and Schlicher, 2005; Gawande and Patil, 2018).

Bioethanol produced from sugar-based feedstocks has the greatest potential of replacing fossil
fuel as transportation fuel due to the low cost of production, low carbon emission and higher
yield in litres of bioethanol per hectare compared to starch-based crops used for production of
bioethanol (de Carvalho et al., 2016; Marzo et al., 2019). For instance, the Brazilian example
of 1G sugarcane-ethanol has extensively demonstrated the environmental and social-economic
benefits of the technology (Crago et al., 2010; de Carvalho et al., 2016). Sugarcane contain
about 12-17 % of soluble sugars in the stem juice of which 90 % are sucrose and 10 % are
glucose and fructose in equal proportions (Nicola et al., 2011). During milling process about
95 % of the soluble sugar is extracted leaving the bagasse that is often used as fuel in a boiler
for cogeneration of electricity and heat (Nicola et al., 2011). The cane juice (sugar) obtained
after extraction undergoes clarification and concentration before fermenting to bioethanol using
the yeast known as Saccharomyces cerevisiae (baker’s or brewer’s yeast) (Naik et al., 2010).
The products from the fermentation broth are separated and distilled to obtain hydrated
bioethanol (approximately 95 %), which can be further dehydrated to obtain 99.99 % bioethanol
using molecular sieve or azeotropic distillation. The technology for 1G bioethanol production

4
Journal Pre-proof
has been in use commercially in US, Brazil and some part of Europe for more than two decades
(Goldemberg, 2007; Taber, 2013).

Several experimental work on the utilization of sugar-based feedstocks for production of 1G


bioethanol have been reported in literature (Table 1). These studies revealed that 1G bioethanol
has been produced using juice extracted from sugar-based feedstocks such as sugarcane, sweet
sorghum, watermelon, sugar beet, Cashew apple. For 1G bioethanol production, the batch mode
fermenter is commonly used for laboratory scale production. For optimum performance of the
microorganism in converting the sugar to bioethanol, parameters such as pH, temperature, and
the speed of agitation must be suitable. For a typical sugar juice fermentation process, pH,
temperature and agitation speed range of 3.7-5.5, 30-40oC, and 150-300 rpm, respectively are
commonly used. The conversion of sugar in the fermenter to bioethanol is to a large extent
dependent on the type of microorganism used. S. cerevisiae is the most popular microorganism
commonly used in 1G fermentation. However, several engineered strain of the S. cerevisiae
such as S. cerevisiae TISTR 5048, S. cerevisiae CICC 1308, S. cerevisiae NP01, S. cerevisiae
Y2084, S. cerevisiae F118 strain have also be used in the 1G sugar conversion to bioethanol.
The comparison of the ethanol yield obtained using the juice extracted from the sugar-based
crops are depicted in Figure 1. The ethanol yields varies with the type of sugar-based feedstock
and the mode of the fermenters used. It is obvious that the use of the un-engineered S. cerevisiae
for the conversion of the sugar obtained from the various juice resulted in the lowest ethanol
yield (Deesuth et al., 2015; Gomez-Flores et al., 2018). However, higher ethanol yields were
obtained using the engineered strain of the S. cerevisiae (Laopaiboon et al., 2009, 2007; Liang
et al., 2008; Liu et al., 2008; Razmovski and Vučurović, 2012; Wijaya et al., 2018).

B
FB B B
100 RB B SB B
90 B B
B
80
70
60 B
Bioethanol yield (%)

B
50
40
30
20
10
0
ice

h u ce

ga et m uice

ice

En hum e

ca ice

ice

ice
s

Sw w a uice
Su er m juic

Sw org juic

rg juic
se
i

ui
ju

ju

ju

ju

ju

ju
ee olas
j

j
J
m

r b elon

Su hick

t s ple

ne

m
et
an

an
u

hu

hu
Ca r be

p
gh

rc

rc
t
g

gy
ga

ga
or

or

or
t

So
ga
e

er
ts

ts

ts
Su

Su
at

sh
rb
ee

ee

ee

ee
W

ga
Sw

Sw

Su

Figure 1: Comparison of bioethanol yield from different 1G sugar-based feedstocks (Note FB:
fed-batch, RB: repeated batch, SB: Sequential batch, B: Batch)

5
Journal Pre-proof

Despite the successes recorded globally in commercializing 1G bioethanol production, it is still


being faced with some constraints such as food versus fuel debate (use of land suitable for food
production for bio-energy crop production; sugar is also considered as human food), high cost
of feedstock, high impact on the environment with the exception of sugarcane bioethanol,
negative impact on biodiversity and competition with scarce water resources in some regions
(Filip et al., 2017; Tokgoz, 2019; Tomei and Helliwell, 2016). These constraints have
stimulated intensive research in the bioethanol production from alternative feedstocks such as
lignocellulosic biomass, also known as 2G bioethanol production (Tan et al., 2008; Akinci et
al., 2008).

6
Table 1: Summary of literatures on 1G bioethanol production from different feedstocks

Feedstock Fermentation Microorganisms Fermenter's Fermentation Total Total Ethanol Productivity Ethanol Reference (s)
mode conditions Time (h) amount of amount concentration (g/L/h) yield
sugar of sugars (g/L) (%)
added to consumed
the in the
fermenter fermenter
Sweet Fed-batch S. cerevisiae pH- 4.8, 108 Reported 234.12 116.62 - 1.08 -1.11 94.12 - Laopaiboon et
sorghum TISTR 5048 Temperature as total g/L 120.28 96.8 al. (2007)
juice (oC)-30, sugar -
Agitation 240 g/L
rate (rpm) -
Static
Sugarcane Repeated S. cerevisiae pH-3.86, 32 Reported 169.4 89.73 2.48 92 Liang et al.
juice batch Temperature as total (2008)
(oC)-30, sugar -
Agitation 173 g/L
rate (rpm) -
Sweet Batch S. cerevisiae pH-4.9, 40-72 Sucrose- Sucrose- 120.68 2.01 99.8 Laopaiboon et
sorghum NP01 Temperature 124.05 124.05 al. (2009)
juice (oC)-30, g/L, g/L,
Agitation Glucose- Glucose-
rate (rpm) - 20.85 g/L, 20.85
Static Fructose- g/L,
16.80 g/L Fructose-
Molasses- 16.80 g/L
78 Molasses-
72.02 g/l

7
Water Batch Dried yeast pH-5.0 36 - 160 Sucrose- Sucrose- 83.2 2.97 80.1 Fish et al. (2009)
melon Temperature 20 g/L, 20 g/L,
juice (oC)-32, Glucose- Glucose-
Agitation 20 g/L, 20 g/L,
rate (rpm) - Fructose- Fructose-
100 36g/L 36g/L
Molasses- Molasses-
210 g/L 130 g/L
Sugar Batch S. cerevisiae pH-5.5, 48 Reported 190 g/L 83.2 2.51 96.56 Razmovski and
beet Temperature as total Vučurović
molasses (oC)-30, sugar - (2012)
Agitation 300 g/L
rate (rpm) -
100
Sugar Batch S. cerevisiae pH-5.5, 48 Reported 294.12 132.39 2.51 96.56 Razmovski and
beet Temperature as total g/L Vučurović
thick (oC)-30, sugar - (2012)
juice Agitation 300 g/L
rate (rpm) -
100
Sugar Batch S. cerevisiae pH-5.0, 72 Sucrose - 182.09 91.16–92.06 1.27 -1.28 84.78– Kawa-Rygielska
beet Juice Temperature 200 g/L, 85.62 et al. (2013)
(oC)-30, Glucose-
Agitation not
rate (rpm) - reported,
150 Fructose-
not
reported
Cashew Batch S. cerevisiae pH-4.5, 120 100 g/L 80.08 74.79 3.11 93.3 Deenanath et al.
apple Y2084 Temperature (2013)
juice (oC)-30,
Agitation
rate (rpm) -
150
8
Sweet Batch S. cerevisiae pH-4.5, 72 220g/L Deesuth et al.
sorghum Temperature- 95.3 2.65 46 (2015)
juice 30oC,
Agitation
time-100rpm
Sweet Sequential S. pH-5.0, 120 227.9 g/L - 109.5 3.17 89.6 Sasaki et al.
sorghum Batch cerevisiae strain Temperature (2017)
juice BY4741 (oC)-30,
Agitation
rate (rpm) -
150
pH-5.0, 72 179.72g/L 10.68 0.46 90 Thammasittirong
Energy Batch S. Temperature- et al. (2017)
cane juice cerevisiaeND48 37oC,
Agitation
speed-
100rpm
Sugarcane Batch Saccharomyces pH-5.0, 72 90 g/L 87.7 g/L 45.6 0.60 55 Gomez-Flores et
juice cerevisiae Temperature al. (2018)
(oC)-30,
Agitation
rate (rpm) -
200
Sorghum Batch Saccharomyces pH-5.2,
juice cerevisiae F118 Temperature- 120 243.84g/L n.r 100.37 3.59 86.19 Wijaya et al.
strain 30oC, (2018)
Agitation
rate (rpm)-
150

9
Journal Pre-proof
2. 2G bioethanol production
Second generation bioethanol production is built on the shortcomings of 1G bioethanol
production by ensuring sustainable use of non-food feedstocks for biofuel production. In this
section, an overview of the use of 2G feedstocks for bioethanol production, the pre-treatment
strategies used for the feedstocks, as well as the saccharification and hydrolysis of the pre-
treated feedstocks are presented.

2.1 Feedstocks for 2G bioethanol production

Lignocellulosic biomass which is the main feedstock for 2G production can be grouped into
agricultural residues (sweet sorghum bagasse, sugarcane bagasse and straw, corn stover) (Goh
et al., 2010), forest residues (Jiang et al., 2019), and energy crops, such as herbaceous or woody
plants (Zabed et al., 2016). Lignin, cellulose and hemicellulose are the main chemical
components of lignocellulosic biomass which can be converted to biofuel through
thermochemical or biological processes (Ma et al., 2019; Shi et al., 2019). The chemical
composition of each of the biomass varies from one another. The chemical composition of
selected agricultural residue-based lignocellulose is shown in Table 2. Except for leaves and
grass, there is higher amount of cellulose in corn stovers, sugarcane bagasse, wheat straw, sweet
sorghum bagasse, and energy cane bagasse (Banerjee et al., 2010). However, the cellulose is
usually interlinked with hemicellulose to form an enclosed complex of cellulose-hemicellulose.
The complex cellulose-hemicellulose if further enclosed by lignin which make it difficult
naturally to unlock the fermentable sugar through hydrolysis.

Table 2: Chemical composition of selected Lignocellulosic biomass

Type of lignocellulose Chemical composition


Cellulose Hemicellulose Lignin Reference
Corn stover 38-40 24-26 7-19 Saini et al.
(2015)
Leaves and grass 15.3 10.5 43.8 Schmitt et al.
(2012)
Sugarcane bagasse 42-48 19-25 20-42 Kim et al.
(2011)
Wheat straw 33-38 26-32 17-19 Saini et al.
(2015)
Sweet sorghum bagasse 34-45 18-27 14-21 Saini et al.
(2015)

2.2 Pre-treatment of 2G feedstocks

Research efforts in 2G bioethanol production have centred on the use of efficient pre-treatment
technology to break the ties between the lignin-cellulose-hemicellulose complexes. An efficient
pre-treatment method releases substantial amounts of cellulose and hemicellulose that can be
hydrolytically converted to fermentable sugar by appropriate enzymes (Rajendran et al., 2018).
Monomeric sugars such as glucose and xylose are produced from the hydrolysis of cellulose
and the hemicellulose portion of the lignocellulosic biomass (Demirbas, 2007). The monomeric
sugars formed during hydrolysis are then converted to bioethanol by fermentative organism
such as yeast (Di Donato et al., 2019). High yields of soluble sugars from lignocellulose
10
Journal Pre-proof
biomass require extensive pre-treatment and high enzyme dosages, both of which add
significant process costs compared to 1G ethanol production; while still producing a more dilute
ethanol product with higher energy demands/costs in distillation, compared to 1G (Chen et al.,
2017; Jönsson and Martín, 2016; Ma et al., 2019). In order to have access to these fermentable
sugars from the lignocellulose materials, it is necessary to disrupt the rigid structure of the plant
biomass, to make it amenable to enzymatic hydrolysis which is the main purpose of pre-
treatment (Chen et al., 2017). Effective pre-treatment is required to liberate the cellulose from
the lignin seal, and to disrupt the crystalline structure of cellulose, to make it accessible for
subsequent enzymatic hydrolysis, allowing completion of the latter with acceptable yields and
process time (productivity) (Kumar and Murthy, 2011). For pre-treatment to be effective,
formation of sugar must be enhanced, degradation of carbohydrate must be avoided, formation
of inhibitory products must also be prevented and the process must be cost effective (Kim et
al., 2019; Carrasco et al., 2010). Lignocellulose pre-treatment can be done using physical,
chemical, physicochemical and biological means. The physical pre-treatment includes
mechanical communition and irradiation (Chen et al., 2017). Chemical pre-treatments which
include, acid pre-treatment (concentrated or diluted), alkaline pre-treatment, ionic liquid pre-
treatment, organosolve, sulphite pre-treatment, alkaline wet oxidation and ozone pre-treatment
has been widely investigated (Chaturvedi and Verma, 2013). The physicochemical pre-
treatments include steam explosion (catalysed and uncatalyzed), liquid hot water pre-treatment,
ammonia fiber explosion and CO2 explosion. The biological means involved the use of micro-
organism for the pre-treatment of lignocellulose. The detail description, merit and the
shortcomings of each of the pre-treatment methods have been extensively reported by Zabed et
al. (2016). The authors concluded that due to the variation in the physicochemical properties
and the chemical compositions of each of the lignocellulosic biomass, there is no stand alone
method for the pre-treatment of lignocellulose biomass. Nevertheless, pre-treatment methods
can be combined in such a way that it targets the particular chemical components in the
lignocelluse.

Amongs all the pre-treatment methods, steam explosion have been widely investigated (Chen
et al., 2017). Steam explosion is a promising pre-treatment method for large scale bioethanol
production and has been used successfully on demonstration and small scale cellulosic
bioethanol plant (Duque et al., 2016). In this method the lignocellulose materials contained in
a reactor is exposed to high pressure saturated steam for a short duration followed by a sudden
depressurization in order to explode the biomass materials (Zhao et al., 2015). The explosive
decomposition lowers the temperature and further quenches the process (Martín et al., 2018)
Steam pre-treatment presents advantages such as high recoveries of hemicellulose, significant
increase of the accessibility of the cellulose to enzymatic hydrolysis, low energy requirements
compared to other pre-treatment options and no recycling or environmental costs involved, as
well as low/none chemicals consumptions (Amores and Ballesteros, 2013). Like most of the
other pre-treatment process, one of the problems associated with steam explosion pre-treatment
of lignocellulosic biomass is the formation of sugar/lignin degradation products which act as
inhibitors to the subsequent hydrolysis-fermentation processes (Phuttaro et al., 2019; Zhao et
al., 2015).The most abundant of all the sugar degradation products are furfural, 5-
hydroxymethylfulfural (HMF), formic acids, phenolic acid, levulinic acid (Dessie et al., 2019;
Jönsson and Martín, 2016). Acetic acid, which is not one of the sugar degradation products is
formed by the hydrolysis of hemicellulose acetyl group, which at high concentration tends to
11
Journal Pre-proof
inhibits the yeast during fermentation (Shuai et al., 2010; Li et al., 2019). The concentration of
the biomass degradation products is proportional to the severity of the pre-treatment process
and varies depending on the types of lignocellulose materials and methods of pre-treatment
(Olofsson et al., 2008). Although, the formation of sugar degradation products has been a major
concern in pre-treatment of lignocellulose materials, the big question is how severe must be the
pre-treatment conditions to have acceptable cellulose digestibility in the subsequent enzymatic
hydrolysis? Several authors have attempted to answer this question through optimization of the
pre-treatment processes by maximising sugar recovery while avoiding unacceptable high
concentrations of inhibitors (Pan et al., 2006; Li et al., 2010; Castro et al., 2014).

Inhibition effects of different degradation products from steam pretreated lignocellulosic


biomass such as wheat straw (Bellido et al., 2011), corn stover (Takada et al., 2019), softwood
(Takada et al., 2019), sugarcane bagasse (Laser et al., 2002; Martín et al., 2007) and sweet
sorghum bagasse (Pengilly et al., 2015) have been extensively investigated. One of the ways
proposed to reduce inhibition effects on the downstream processing is integration of sugar-rich
1G stream with the 2G stream. This will serve the purpose of diluting the inhibitors thereby
reducing its effects. Rohowsky et al. (2013) demonstrated this concept by the addition of sugar-
juice to hydrothermally pretreated sweet sorghum bagasse and concluded that co-fermentation
of juice with 2G sugar stream can help reduce inhibition effects and further increase ethanol
concentrations in the fermentation broth. Different techniques for pre-treatment and
hydrolysing lignocellulose materials, as well as process configurations, have been investigated
with the intention of mitigating the low bioethanol productivity and reduce cost in stand-alone
2G bioethanol production process (Conde-Mejía et al., 2012; Conde-Mejía et al., 2013).Steam
explosion pre-treatment can be used with different configurations of hydrolysis and
fermentation steps. Complete acid hydrolysis is not often considered as a result of issue with
acid recovery.

2.3 Hydrolysis and Fermentation

The pre-treatment stage is often followed by hydrolysis and fermentation stage. Where
enzymatic hydrolysis is used, there is the possibility of different level of process integrations
such as separate, consolidated bioprocessing (CBP), separate hydrolysis and fermentation
(SHF) and simultaneous saccharification, and fermentation (SSF), simultaneous
saccharification and co-fermentation (Conde-Mejía et al., 2012). The successful utilization of
each of the configuration for hydrolysis and fermentation is dependent on the chemical
components and the right proportionate ratio of each of the components used in each stage
(Ojeda et al., 2011). The utilization of the SHF for solid hydrolysis and sugar fermentation
entails performing the hydrolysis of the pre-treated solid hydrolysis and the fermentation of the
sugar released during the hydrolysis one at a time in separate vessel (Loaces et al., 2017). The
optimal conditions of the parameters (pH and temperature) for both the solid hydrolysis and the
sugar fermentation can be achieved (Cotana et al., 2015). Nevertheless, one major constraint of
the SHF is the inhibition of the activity of the cellulases by the frequent accumulation of sugars
(glucose and cellobiose) (Loaces et al., 2017). The high cost of running the separate two
separate stages is also a major constraint (Kadhum et al., 2017). On the other hand, in SSF both
the solid enzymatic hydrolysis and the sugar fermentation are run concurrently in the same
vessel (Loaces et al., 2017). This implies that the sugar (glucose) produced from the solid

12
Journal Pre-proof
hydrolysis is immediate converted by the fermenting microorganism to bioethanol. The use of
SSF has the advantages of overcoming the inhibition of the activity of the cellulases since there
is continuous utilization of the released sugar (Mithra et al., 2018). In addition, the
contamination of the microbes in prevented due to the presence of ethanol in culture broth
(Dahnum et al., 2015). Compared to SHF, the rate of hydrolysis is faster in SSF which
invariable leads to higher ethanol productivity and yield (Szambelan et al., 2018). However,
one major challenges with SSF is the inability of the fermentation organism to utilize the
pentose sugar (Loaces et al., 2017). This has led to the discovery of fermentation organism that
can utilize both pentose and hexose sugars for ethanol production often refer to as co-
fermentation. Therefore, in SSCF the enzymatic hydrolysis and the co-fermentation of the
pentose and hexose sugars occurs in the same vessel concurrently and invariable prevent
creating a separate stage for the fermentation of the pentose sugar or utilizing it for biogas
production through bio-digestion (Pandey et al., 2019).

Amongst these process options, SSF and SSCF of steam pre-treated lignocellulose biomass are
the most promising approaches for 2G bioethanol production which can be adopted for
integration with 1G bioethanol production process. This is due to the several advantages that
SSF and SSCF present over SHF. In SSF and SSCF, the number of process stages is reduced
compared to SHF, where separate process equipment is needed for the hydrolysis and the
fermentation steps which could also lead to reduction in investment cost. The problem of end
product inhibition of enzymatic hydrolysis, which is caused by accumulation of glucose in SHF,
is reduced in SSF because both the hydrolysis and fermentation steps are consolidated in one
vessel enabling the immediate fermentation of the sugars formed from hydrolysis (Olofsson et
al., 2008). The latter will reduce the overall enzyme dosages required for hydrolysis of pre-
treated lignocellulose. However, the major challenge with the SSF and SSCF layout is the
determination of the optimum working conditions for both the enzyme and the yeast. Choosing
a working temperature in favour of the fermentation organism results in a process temperature
that is too low for efficient working of the enzymes (Alfani et al., 2000).

The difficulty of separating the yeast from lignin after fermentation for re-use is also one of the
constraints of using SSF and SSCF (Raud et al., 2019; Sewsynker-Sukai and Gueguim Kana,
2018). This usually requires the production of fresh yeast for every fermentation run, rather
than re-use of yeast which represents a yield loss if the yeast is produced from the carbohydrate
within the process or increase the running cost if it is supplied externally (Mohd Azhar et al.,
2017). It is easier to recycle and re-use yeast during SHF because the solubilized sugar is
separated from the solids, before being fed to the fermenter (Maslova et al., 2019). The
fermentation product is free of solid plant biomass, and the yeast can easily be recycled from
this product for re-use in fermentation which reduces the amount of the lignocellulose-derived
sugars required for yeast production (Sudiyani et al., 2019). The required production of fresh
yeast in SSF can be avoided if there is an efficient way of separating the yeast from the
fermentation product for re-use. The choice between SSF and SHF for stand-alone 2G ethanol
production, considering their respective advantages and disadvantages in terms of their impacts
on the downstream processing, has been investigated (Ask et al., 2012; Dahnum et al., 2015;
Loaces et al., 2017; Mithra et al., 2018; Wirawan et al., 2012).

13
Journal Pre-proof
In order to maximise ethanol yield from cellulosic bioethanol production, it is necessary to
consider the utilization of xylose sugars in the hemicellulose portion of lignocellulose while
reducing inhibitory effects on the fermentation process as much as possible (Liu et al., 2019).
Since the commonly used Saccharomyces cerevisiae has been shown not to be effective in
fermenting xylose sugars, intensive research has resulted into development of engineered
strains of S. cerevisiae that can effectively co-ferment both glucose and xylose sugars as shown
in Table 3. A situation where hydrolysis-fermentation of the cellulose takes place
simultaneously with the co-fermentation of xylose to ethanol is known as simultaneous
saccharification and co-fermentation (SSCF) (Li et al., 2019). Co-fermentation of glucose and
xylose sugars derived from wheat straw to bioethanol has been successfully demonstrated on
an industrial scale production of 2G bioethanol by DONG Energy and Royal Dutch DSM (
Europian Biofuel Technology Platform, 2013). This co-fermentation process was reported to
give 40 % increase in ethanol yield per ton of straw compared to fermentation of only glucose
sugars to ethanol. This also led to significant reduction in the production cost of 2G bio-ethanol
since yield of ethanol is maximized from the feedstock. Stand-alone 2G bioethanol production
from co-fermentation of xylose and glucose sugars using SSCF approach has been extensively
investigated as shown in Table 1 while it has only been considered for integrated 1G2G
bioethanol production from whole wheat by Erdei et al. (2013).

Bioethanol productions from different cellulosic feedstocks in the stand-alone 2G process


configuration using different types of pre-treatment methods, microorganisms and varying
conditions in SSCF of xylose and glucose sugars are shown in Table 3. Co-fermentation of
xylose and glucose sugars using SSCF can be done using batch or fed-batch mode. Also, the
pre-treated lignocellulosic materials can undergo pre-saccharification prior to SSCF based on
investigations of Krishnam et al. (2008), Lau at al. (2010) and Souza et al. (2012). Conducting
SSCF of xylose and glucose in fed-batch mode rather than batch mode helps to overcome
limitations such as high viscosity and problem of mass transfer in the fermenter. This invariably
tends to increase the ethanol concentrations and ethanol productivity. Similarly, subjecting pre-
treated lignocellulosic biomass to presaccharification prior to SSCF of xylose and glucose has
proven to increase the ethanol concentrations and productivity. Souza et al. (2012) established
this fact in their investigation of the influence of presaccharification, fermentation temperature
and yeast strain on ethanol production from sugarcane bagasse. Two different thermo-tolerant
microorganisms namely S. cerevisiae LBM-1 and Kluyveromyces marxianus UFV-3 were used
for co-fermentation of xylose and glucose at a temperature of 50 oC during pre-saccharification.
The result shows that the ethanol productivity increases with pre-saccharification time. The
authors concluded that using thermo-tolerant fermentative organisms and pre-saccharification
stage significantly increase ethanol yield, concentrations and productivity. The major challenge
of this process is the inefficient co-fermentation of xylose as a result of higher glucose
concentrations in the medium. As shown in Table 1, ammonia fiber explosion pre-treatment
seems to be more effective in releasing the fermentable sugars compared to other means of pre-
treatment. In comparison with other types of pre-treatment, the pre-treatment of corn stover
using ammonia fiber explosion resulted in the highest ethanol yield, ethanol concentration and
productivity of 93 %, 47.5 g/L and 0.77 g/L. h, respectively. The high amount of ethanol yield,
concentration and productivity obtained using ammonia fiber explosion could be attributed the
possibility of a high fermentable sugar yield, and low formation of sugar degrading products
(Badiei et al., 2014). However, Baral and Shah (2017) in their techno-economic studies of steam
14
Journal Pre-proof
pre-treatment, dilute sulfuric acid pre-treatment, ammonia fiber explosion pre-treatment and
biological pre-treatment of corn stover revealed that ammonia fiber explosion pre-treatment has
a high production cost compare with steam pre-treatment, dilute sulfuric acid pre-treatment but
lower than the biological pre-treatment. Besides the types of pre-treatment methods, using the
appropriate conditions in the fermenter suitable for the conversion of the sugar to ethanol by
the microorganism is of paramount important. Study have shown that for optimum performance
of the microorganism during SSCF of the sugar, the pH of the medium and the temperature of
the fermenter must be in the range of 4-5 and 30-50 oC respectively. The type of fermenter and
their configuration also have significant effects on the conversion of sugar to bioethanol.
Olofsson et al. (2008) employed batch and fed-batch fermenter for the SSCF of sugar obtained
from dilute H2SO4 catalyzed steam explosion pre-treatment of wheat straw. Their work revealed
that a higher ethanol yield, concentration and productivity were obtained using the Fed-batch
SSCF configuration. Also, in a similar study, Carrasco et al. (2010) investigated the effect of
using batch and fed-batch fermenter for the SSCF of sugar obtained from SO2 catalyzed steam
explosion pre-treatment of braze straw. Their study also confirm that a higher ethanol yield,
concentration and productivity were obtained using the fed-batch SSCF configuration. It is
certain that the bioethanol yield obtained from various 2G feedstocks as shown in Figure 1 is
dependent on how the efficiency of the pre-treatment technique, the saccharification process
and the hydrolysis strategies.

100
90
80
70
60
Bioethanol yield (%)

50
40
30
20
10
0
se

an traw

af

a)

se

te

od
va
ve
ss

W gass
ra

ra
le

av

as
s

as

w
ra
ga

ga

to
st

st

w
Br
ne

rd
s

B
S
ba

ba

ba

ba
at

at

at

Ha
na
a

ja

a
rn
rc
he

he

he
aj
ne

e
a

na
Co

an

an
ga

(P

(P
W

W
a

Ba
rc

rc

Su

rc

rc
w

w
ga

ga

ga

ga
a

ra
r
st

st
Su

Su

Su

Su
e

e
av

av
Br

Br

Figure 2: Comparison of bioethanol yield obtained from different 2G feedstocks

15
Table 3: 2G bioethanol production from different feedstocks using different pre-treatment method, microorganism and fermentation conditions

Feedstock Pre-treatment Microorganism Conditions in the Conversion Ethanol Ethanol Ethanol Reference
fermenter process yield Concentration Productivity
(%)* (g/L) (g/L.h)
Sugarcane SO2(2 % w/w Saccharomyces pH-5.0 , Temp.-32 5-7.5 % 59 26.70 0.27 Rudolf et al.
bagasse moisture) cerevisiae oC WIS/Batch/SSCF (2007)
catalyzed steam TMB3400,
explosion pre-
treatment
Wheat Dilute H2SO4 Saccharomyces pH-5.0 , Temp.-34 9% 59 33.20 0.33 Olofsson et
straw catalyzed steam cerevisiae oC WIS/Batch/SSCF al. (2008)
explosion pre- TMB3400,
treatment
Wheat Dilute H2SO4 Saccharomyces pH-5.0 , Temp.-34 6-9 % WIS/Fed 71 38.10 0.38 Olofsson et
straw catalyzed steam cerevisiae oC batch/SSCF al. (2008)
explosion pre- TMB3400,
treatment
Sugarcane Ammonia Fibre Saccharomyces pH-4.8 , Temp.-30 18 % 92 33.70 0.28 Krishnan et
bagasse explosion pre- cerevisiae oC WIS/batch/PSSCF al. (2010)
treatment 424A LNH-ST
Sugarcane Ammonia Fibre Saccharomyces pH-4.8 , 18 % 92 36.70 0.51 Krishnan et
leaf explosion pre- cerevisiae Presaccharification WIS/batch/PSSCF al. (2010)
treatment 424A LNH-ST Temp.-50 oC,
Fermentation
Temp.-30 oC
Corn Ammonia Fibre Saccharomyces pH-5.5 , 18 % 93 47.50 0.77 Lau & Dale
Stover explosion pre- cerevisiae Presaccharification WIS/batch/PSSCF (2010)
treatment AX101 Temp.-50 oC
Fermentation
Temp.-30 oC
Brave straw SO2(2% w/w Saccharomyces pH-5.5 , 7% 80 32.50 0.71 Carrasco et
(Paja Brava) moisture) cerevisiae Presaccharification WIS/batch/SSCF al. (2010)
catalyzed steam TMB3400, Temp.-50 oC
explosion pre- Fermentation
treatment Temp.-30 oC
16
Brave straw SO2(2% w/w Saccharomyces pH-5.5 , Temp.- 6-9 % WIS/Fed 81 40.40 0.85 Carrasco et
(Paja Brava) moisture) cerevisiae 30 oC batch/SSCF al. (2011)
catalyzed steam TMB3400,
explosion pre-
treatment
Sugarcane Dilute acid Kluyveromyces pH-4.8 , 8 % WIS/ 45 14.36 1.79 de Souza et
bagasse pretreatment marxianus Presaccharification Batch/PSSF al. (2012)
UFV-3 Temp.-50 oC
Fermentation Temp-
42oC
Wheat Dilute H2SO4 Saccharomyces pH-5.0 , Temp.- 8-11 % WIS/Fed 69 38.00 0.38 Olofsson et
straw catalyzed steam cerevisiae 30 oC batch/SSCF al. (2013)
explosion pre- TMB3400,
treatment

Sugarcane Alkaline Cellulase pH-4.5, Temp-34 oC 24 % WIS Fed- 84 72.04 1.80 Ye et al.
bagasse pretreatment batch/SSF (2018)

Banana Acid-catalysed Saccharomyces pH-4.8, Temp-38 oC 17.6 % WIS PSSF 85 42.00 0.40 Guerrero et
waste steam explosion cerevisiae al. (2018)
Ethanol Red
Hardwod Hydroperoxide Saccharomyces pH-5, Temp-37 oC 5 %WIS SSF 80.7 14.8 0.32 Song et al.
acetic acid cerevisiae, (2019)
pretreatment Pichia stipites Subsamran
Vetiver Alkaline-acid Saccharomyces pH-5.0, Temp-30 oC 5 %WIS SSF 3.70 et al. (2019)
grass pretreatment cerevisiae
TISTR 5339
*Ethanol yield was based on fermentable sugar

17
Journal Pre-proof

3.0 Integration approach for 1G2G bioethanol production

Integration of 1G2G bioethanol production could provide solutions to some of the challenges
in standalone 2G processes, which could eventually lead to breakthrough in development of
viable technology for bioethanol production. According to van Zyl et al. (2011) “the cost
disadvantage of standalone second-generation biofuels processes may be addressed through
innovative methods of process integration, which will help to minimize the capital investment,
maximize energy efficiency and improve overall economics”. Integration of 1G and 2G
bioethanol production will provide the opportunity to utilise the whole agricultural crops for
ethanol production, which will increase the yield of ethanol per hectare of cultivated land for
feedstocks such as sugarcane, sweet sorghum, corn, wheat, sugar beet etc. This will also
facilitate the introduction of cellulosic materials in the production of bioethanol at industrial
scale (Erdei et al., 2013a). It is important to optimize cellulose conversion of the feedstock to
bioethanol through appropriate handling and utilization of all process streams. This is essential
because of high costs of the feedstocks which accounts for more than 18 % in the case of
lignocellulosic biomass (Banerjee et al., 2010).

3.1 1G2G integration strategies

The benefits obtained during integration of 1G2G bioethanol productions depends on the point
of integrations such as energy (or thermal) integration, integration around CHP, distillation
column and fermentation step. Energy (particularly thermal) integration between streams in the
1G2G processes using Pinch technology can be done to increase the overall energy efficiency
of the combined process (Cardona and Sánchez, 2007). Studies have shown that thermal
integration enhances ethanol production through reduction in steam consumption in the
integrated process (Palacious-Berech et al., 2011). Dias et al. (2009) used Pinch technology
for optimization of energy consumption in order to estimate the amount of bagasse that will be
available for 2G ethanol production rather than burning all the bagasse in a boiler for steam
and electricity production in 1G process. The author concluded that thermal integration using
Pinch technology could increase the availability of bagasse in 2G ethanol production process
while reducing the demand for hot utilities. Similarly, Dias et al. (2011b) investigated how
thermal integration can be used to improve bioethanol production from sugarcane. The authors
used Pinch point method to evaluate the streams of the integrated 1G and 2G-SHF processes
which were available for thermal integration. The process streams were represented using
enthalpy-temperature diagrams to determine the streams’ thermal integration target. The hot
and cold streams parameters for both conventional and double-effects distillation column were
considered in the thermal integration analysis of the ethanol production. The results show that
the thermal integration reduces steam consumption in the anhydrous bioethanol production
with double-effects distillation column having the lowest steam consumption. The author
concluded that thermal integration provides significant energy savings in the bioethanol
production process especially with double-effect distillation column, thus allowing for
availability of more bagasse for ethanol production and increase in surplus electricity. Contrary

18
Journal Pre-proof

to the submission of Dias et al. (2011b), double effect-distillation column requires a more
complex control system and high capital cost.

The integration around CHP is done to produce steam and electricity for the 1G2G processes,
as well as surplus electricity for export to the grid. Production of electricity can either be done
using conventional power generation plants or CHP plants. Convectional power generation
plants have the limitations of low efficiency (about 40 %) and non-utilization of the waste heat
produced. CHP plant, on the other hand can attain plant efficiency of about 90 % by utilizing
the waste heat produced for process heating (Magnusson, 2006). Dias et al. (2013c) reported a
modeling work on cogeneration in integrated 1G2G bioethanol production from sugarcane
using SHF for solids hydrolysis-fermentation. Aspen Plus was applied to compare different
configurations of the CHP for steam productions at 22, 42, 65 and 82 bars with bagasse, and
biogas produced from pentose biodigestion used as fuels in the CHP. The different
configurations of the integrated 1G2G bioethanol production with the cogeneration systems
were compared based on economic and environmental life cycle analysis. Economic analysis
by the authors revealed that co-generation of electricity and steam integrated with 1G2G
bioethanol production enhances the profitability of the integrated process, even in relative small
amounts of electricity production compared to the same scenario without CHP. The
configuration however has higher environmental impacts than the 1G process based on life
cycle assessment. The authors concluded that integrated 1G and 2G-SHF bioethanol production
with CHP systems presents better economics performance than an integrated process without
CHP.
The integration at the distillation step entails using one common distillation column for the
product streams from separate 1G and 2G fermentation steps rather than using two separate
distillation columns. Integrating at the distillation step is easier in a situation where an existing
1G bioethanol production plant is to be combined with a new 2G plant (Arvinius et al., 2010).
High ethanol concentration is not easily obtained in 2G fermentation as a result of low sugar
concentrations, which is caused by difficulties in handling high substrates loading. The
difficulties are caused by high substrate viscosity, insufficient mass transfer and substrates
inhibition which lead to reduction in the efficiency of the enzymatic hydrolysis and
fermentation, thus producing lower ethanol yield and concentrations during fermentation. This
can be circumvented by using high enzyme dosage which will lead to increase in production
cost. On the contrary, high ethanol concentration and yield can easily be attained during 1G
fermentation process due to absence of sugar degradation products that can inhibit the
fermentation process and the presence of high concentration of sugar. The energy demand for
distillation is low in 1G process due to high concentration of ethanol (about 80 – 120 g/L) at
the fermentation broth as shown in Table 4. Economic synergy could be maximized by
distilling the combined fermentation broths from 1G and 2G processes, since mixing the two
streams will give ethanol concentration higher than the required minimum 40 g/L to make the
process economically feasible.

Macrelli et al. (2014) considered the option of integrating 1G and 2G-SHF ethanol production
at the distillation step. They investigated the effect of production and market factors on ethanol
profitability on an integrated 1G and 2G-SHF bioethanol productions from sugarcane Two

19
Journal Pre-proof

options namely design option (type of feedstock and pentose utilization) and design variables
(Enzymatic hydrolysis (EH) residence time and WIS loading in EH) were considered with
respect to production while enzymatic hydrolysis efficiency (EHE), electricity selling price,
whole sale ethanol selling price, sugarcane cost and leave cost were the cost factors evaluated.
The 1G and 2G bioethanol production was integrated at the distillation step and modelled using
Aspen Plus simulation software with data obtained from the literature. Twelve different
scenarios were evaluated based on design options (type of feedstocks) and design variables
(EH residence time and WIS loading in EH). The authors assumed 95 % yield in each
fundamental conversion step (pre-treatment, EH, fermentation, solid/liquid separation,
anaerobic digestion). The result of the findings shows that cost of enzymes, addition of leaves,
WIS loading, EH residence time had smaller effect compared to the cost of additional filter
press and tank needed for the anaerobic digestion of pentose as well as larger CHP which is
required due to co-combustion of a higher amount of biogas in addition to EH solid residue.
The total investment cost was higher in integrated 1G and 2G processes compared to stand
alone 1G and 2G respectively. More so, the minimum ethanol selling price (MESP) was lowest
in the integrated 1G and 2G process with bagasse and leaves as feedstocks (Figure 1). The
authors concluded that integrated 1G and 2G bioethanol productions from combination of juice,
bagasse and leaves with pentose utilization for bioethanol is more economically attractive
compared to integrated 1G and 2G with only bagasse as feedstock and pentose utilization for
biogas and standalone 1G and 2G bioethanol production. The authors did not however consider
the limitations of integrating at distillation step. Integrating at the distillation step, rather than
fermentation step, increases the number of process stages which in turn increase the investment
cost. Also, the problem of end-product inhibition which is peculiar to 2G process using SHF
could have been reduced if integration was done at the fermentation step using SSF for solids
hydrolysis-fermentation. The assumptions made in the study needs to be experimentally
validated.

Integration of 1G and 2G bioethanol production using SSF for ethanol production from
cellulose has only been considered to a limited extent in the literature using wheat and sweet
sorghum as shown in Table 4. Arvinius et al. (2010) and Erdei et al. (2013) investigated SSF
layout using whole wheat as feedstock. The work by Arvinius et al. (2010) focused on
integrating 2G bioethanol production from wheat straw into an existing 1G bioethanol
production from wheat at the distillation step with an option of biogas production. The wheat
straw was impregnated with dilute sulphuric acid and pre-treated with steam explosion at 190
oC for 10 mins. SSF was used for the solids hydrolysis-fermentation. The authors removed 50

% of the liquid from the pre-treated slurry using a filter press before being fed into the SSF
using S.cerevisiae as the fermentation organism The filtered liquid was mixed with stillage for
biogas production. The research findings show that the integrated 1G and 2G bioethanol
production was better and has more economical prospect than the stand alone 1G and 2G
processes. However, the energy demand was found to be higher than the stand alone 1G and
2G processes. Also, the author did not consider the implication of reducing inhibition by
washing the pre-treated slurry with water. This might lead to loss of potential fermentable
sugars in the hydrolysate and also increase the processing cost due to the large volume of water
required and the need for water treatment (Lu et al., 2013). Integration of 1G and 2G can be

20
Journal Pre-proof

done at the fermentation or distillation step. Doing the integration at the distillation step is
easier in a situation where an existing 1G bioethanol production plant is to be combined with
a new 2G plant. In this case, only the distillation unit, downstream process and CHP unit will
be needed to be built. The second option which is more advantageous is to integrate the 1G and
2G streams in the fermentation stage which requires substantial redesigning of the existing
process when compared with the former. This has the merit of reducing the effect of the
inhibitors associated with 2G through combination with sugar-rich 1G stream and also ensure
that biological benefits are maximized. But appropriate feeding mechanism using fed-batch is
necessary for this to be efficient.

Erdei et al. (2010) investigated the effect of mixtures of wheat straw and whole meal on
bioethanol production. The wheat straw was impregnated with a weak sulphuric acid and steam
pre-treated while the wheat meal undergoes pre-saccharification. The steam pre-treated wheat
straw (SPWS), the pre-saccharified wheal meal (PWM) and various mixtures of both SPWS
and PWM were used as substrate in SSF using a total water insoluble solid (WIS) content of 5
%. Five PWM: SPWS WIS ratios for the mixtures were investigated. The SSF experiment was
performed using Celluclast 1.5 L and Novozym 188 enzymes at a dosage of 15 FPU/g glucan
and S. cerevisiae. The results of the study revealed that pure SPWS has 68 % ethanol yield
compared to pure PWM which was 91 %. Also, the SSF of the mixtures at different ratios
shows that ethanol yield increased with increase in proportion of PWM. Bioethanol yield of 99
% was obtained when equal ratios of SPWS and PWM were mixed in SSF compared to pure
SPWS and PWM. The authors concluded that a mixed substrate is favourable in terms of final
ethanol yield. The authors did not however consider the utilization of xylose sugar which could
have increased the bioethanol yield and bioethanol concentration. This could have been
achieved using an engineered strain of S. cerevisiae that can co-ferment both glucose and
xylose sugar. In this present study, the effect of varying mixing ratios of sugarcane juice and
steam pre-treatment lignocellulose on bioethanol yield and bioethanol concentration will be
investigated

Integration around the fermentation steps involves joint fermentation of 1G sugar-rich stream
with 2G sugar-streams. This can be achieved by adding juice (syrup) to the 2G-SSF process
(either glucose fermentation only or co-fermentation of xylose and glucose) in such a way that
the glucose concentrations in the fermentation broths will not exceeds 150 g/L, high enough to
cause end-product inhibition of the activities of the enzyme and yeast (Wu et al., 2011). The
feeding of the juice (syrup) into the fermenter can be controlled using fed-batch, i.e. feeding
the syrup (assuming a liquid stream) at a rate lower than the maximum glucose consumption
rate of the organism, yet at a rate greater than the critical rate where the Crabtree effect (a
situation where cellular respiration is inhibited by anaerobically ethanol production by yeast in
the presence of high external glucose concentration) kicks in. Also, fed-batch strategy is
adopted to optimise substrates loading in order to enhance fermentation and to prevent viscosity
and mass transfer limitations. The addition of substrates by pulses during fed-batch feeding
strategy enables the yeast to detoxify some of the sugar degradation products in stepwise
thereby keeping the concentrations of the these compounds low in the fermentation broth
(Xiros et al., 2014).

21
Journal Pre-proof

Feeding of juice into the fermenter for the purpose of integration of 1G sugar stream with SSF
of 2G bioethanol production can be done using a computer controlled peristaltic pump (Ding
and Tan, 2006). This can be achieved by applying different fed batch strategies such as pulse
fed-batch, constant feed rate fed batch, constant residual glucose concentrations fed batch and
exponential fed batch. (Ding & Tan, 2006; Liu et al., 2014). Pulse fed batch strategy entails
feeding the substrate (juice) into the fermenter at varying pulse feeding time in order to achieve
residual sugar (glucose) concentrations less than 5 g/L. Substrate feeding in constant feed rate
fed batch is done at specified feed rate interval that will keep the sugar concentrations in the
fermentation broth between 0 and 5 g/L in order to prevent end product inhibition. The residual
sugar (glucose) in the fermentation broth is maintained at specified concentrations during
constant residual glucose concentration fed batch when the substrate is pumped into the
fermenter. In exponential fed batch strategy, cells are grown at a specific growth rate by feeding
the nutrients into the fermenter at an exponential increase in time.

Erdei et al. (2013) used the whole wheat for 1G and 2G-SSF ethanol production integrated at
the fermentation step. The 1G stream was made of starch from wheat meal while the 2G stream
was made of wheat straw. The wheat straw was steam pre-treated with impregnation of dilute
sulphuric acid at 190 oC for 10 min while the wheat meal was enzymatically hydrolysed in a
flask at 35 oC for 24h. Saccharified wheat meal (SWM) and fermented wheat meal (FWM)
were simultaneously saccharified and fermented with steam pre-treated wheat straw in both
batch and fed-batch mode. The result of the studies shows that the addition of SWM to SSF
steam-pre-treated wheat straw, using S. cerevisiae resulted to ethanol yields of about 90 % of
the theoretical. The findings also revealed that the addition of FWM in a batch mode was toxic
to the S. cerevisiae as a result of the bioethanol content of the FWM .This resulted to very low
yield and high accumulation of glucose compared to low accumulation and fairly high
bioethanol concentration obtained in fed batch mode. One of the constraints with integrating
1G and 2G bioethanol production from starch-based feed stocks such as wheat is the
complexity of the whole process. Separately fermenting and pre-saccharification of the wheat
meal before integrating with the 2G stream will invariable increase the number of process
stream which in turn lead to increase in capital and cost of production.

Both integrations at the distillation and fermentation steps have the advantages of attaining
higher ethanol yield, ethanol concentration and productivity, as well as lower energy demand
compare to stand-alone 1G and 2G processes as shown in Tables 2 and 3. However, integrating
at the fermentation step offers better economic advantages due to reduced number of stages
compare to integrating at distillation step. It also has the merit of reducing the effect of
inhibitors in 2G streams through combination with sugar-rich 1G stream. Appropriate feeding
mechanism using fed-batch is necessary for this to be efficient (Erdei et al., 2010).

Integration at the distillation stage by merely combining the fermentation products from 1G
and 2G will not provide opportunity for full maximization of biological benefits that would
have been obtained from combining the 1G and 2G sugar streams. This concept was proven by
Rohowsky et al. (2013) when investigating the feasibility of bioethanol production from whole
sweet sorghum crop (Juice and bagasse) using SSF for solids hydrolysis-fermentation. In their
investigation, the authors carried out SSF of pre-treated sweet sorghum bagasse and Juice

22
Journal Pre-proof

fermentation separately. Also, another experiment was performed by combining 1G and 2G


sugar streams for joint fermentation. The products from the separate fermentation broths were
distilled to obtain bioethanol concentrations of 33 g/L compared to 44 g/L bioethanol obtained
when the two streams were jointly fermented. The results show that the bioethanol
concentrations obtained during separate fermentation was 42 % lower than the corresponding
concentrations obtained in the combined fermentation. The authors concluded that, the
combined fermentation process offers more biological benefits than separate fermentation of
the 1G and 2G sugar streams.

Macrelli et al. (2012) investigated the effects of integration of 1G and 2G-SHF ethanol
production from sugarcane at the fermentation and distillation steps. Fourteen different process
scenarios for integrating 1G and 2G-SHF ethanol production were considered. The reference
scenario was an autonomous distillery representing the current technology applied in the 1G
bioethanol production. Flow sheets were first design and simulated for the fourteen process
scenarios using Aspen Plus v 7.1. The simulation data used were obtained experimentally from
the previous work of the authors. Capital cost evaluation and investment cost analysis were
done using Aspen Process Economic Analyser. The profitability of the various process
scenarios studied was expressed as MESP for 2G ethanol. The result show that the scenario
integrated at the fermentation step has the highest MESP. The authors also observed that the
combined 1G and 2G-SHF ethanol production was economically feasible for all the scenario
simulated when the average ethanol selling price was 0.53US$/L, despite lower internal return
on investment (IRR< 21.5 %). It was also observed that the addition of leaves almost doubled
the amount of ethanol in 2G ethanol. Electricity price and enzyme cost were found to have
significant effect on the 2G ethanol production cost if part of the leaves and bagasse are not
considered as fuel in boiler for combine heat and power cogeneration. The study did not
consider the effect of variation in the sugar stream from both 1G and 2G on the integrated
process. In addition, the effect of steam explosion pre-treatment on the downstream processing
viz a viz formations of sugar degradation products, which often inhibits the process, were not
considered. Similarly, since SHF approach was used for the 2G ethanol production from
cellulose, overall process might be constrained with the problem of glucose accumulation
which often inhibit the process. This can be overcome by considering the application of SSF
approach for the solid hydrolysis and fermentation.

There remains considerable potential for integration of fermentation and distillation steps in
the 1G and 2G processes for feedstocks such as sugarcane and sweet sorghum. This will
eventually reduce the capital and operating costs as well as ensuring the optimum production
of other essential co-products (Macrelli et al., 2012). The goal of making 2G bioethanol
production more competitive can be realised by integrating the 1G and 2G technologies in a
plant sharing energy and material streams in unit operations for exploiting synergies. The
integrated 1G2G processes will provide opportunity for utilization of the whole crop for
production of bioethanol which has significant benefits in terms of the economic feasibility of
2G ethanol (Dias et al., 2013a and Macrelli et al., 2014). The problem of end-product
inhibition associated with 2G process can be reduced by integration with 1G sugar stream,
which serves the purpose of diluting the inhibition effects (Rohowsky et al., 2013). As a

23
Journal Pre-proof

proposal for further study, the 1G and 2G processes can be integrated at the fermentation step
as shown in Figure 3.

Sugarcane Milling Juice


Steam

Bagasse Juice treatment


CHP Electricity
To Grid
Steam explosion Pre- Clarification
treatment Lignin
CO2 Continuous
Pre-treated liquor feeding

Pre-treated slurry SSF


(Pressed or whole slurry) Fed-batch Distillation

Xylose Enzyme
fermenting Bioethanol
Yeast

Figure 3: Process flow diagram of integration approach as a proposal for further study

24
Journal Pre-proof

Table 4: Summary of literature on experimental work related to integrated 1G2G bioethanol production

Feedstock Pre- Inhibition Solid Microorganism Enzyme Feeding Conditions in 1G-sugar 2G-sugar Ethanol Ethanol Ethanol Reference
treatment concentration loading Dosage strategy the fermenter concentration concentration yield Concentrat Productivity
(g/L) (%) (%) ion (g/L) (g/L.h)
Sweet Liquid hot Acetic acid- 7.50 Saccharomyces 20 FPU/ g Batch/SSF pH-4.8 Total sugar-74 Glucose-36.2 86.00 53 0.74 Rohowsky et al.
Sorghum water pre- 3.4, Formic cerevisiae biomass Temperature- g/L (glucose, g/L Xylose-15 (2013)
treatment acid- 0.9, (Thermosacc) 37oC fructose) g/L
HMF-0.3,
Furfural-1.3
Sugarcane Low Not reported Not Saccharomyces 20 FPU/ g Batch/SHF pH-4.8 225.9 g 91.8 g 88.00 44 3.77 Wu et al. (2011)
Temperatur specify cerevisiae biomass Temperature-
e Alkaline NBRC 0224 30oC
Pre-
treatment
Sweet Alkaline Not reported 3.00 Saccharomyces 30 FPU/g Batch/SHF pH-4.8 125.5 g/L 37 g/L 68.40 58.5 1.21 Kim et al. (2012)
Sorghum pre- cerevisiae biomass Temperature- (Sucrose-25
(juice, treatment 37oC g/L, Fructose-
bagasse & 45.5 g/L,
grain) Glucose-55
g/L)
Wheat Dilute Furfural-2.6 7.50 Saccharomyces 20 FPU/ g Batch/SSF pH-5.0 52.2 g/L 149.1 g/L ( 75.00 59.5 0.98 Erdei et al. (2013)
H2SO4- HMF-0.2, cerevisiae biomass Temperature- (Mainly glucose-122.7
impregnate Acetic acid- 35oC glucose) g/L and
d Steam 2.4, Formic xylose-26.4)
explosion acid-0.4
pre-
treatment
Wheat Dilute Furfural-2.6 8.80 Saccharomyces 20 FPU/ g Batch/SSCF pH-5.0 60.9 g/L 154.5 g/L ( 72.00 56.8 1.86 Erdei et al. (2013)
H2SO4- HMF-0.2, cerevisiae biomass Temperature- (Mainly glucose-123.5
impregnate Acetic acid- KE6-12 35oC glucose) g/L and
d Steam 2.4, Formic xylose-31)
explosion acid-0.4
pre-
treatment
Wheat Dilute Furfural-2.6 7.50 Saccharomyces 20 FPU/ g Fed- pH-5.0 52.2 g/L 149.1 g/L ( 76.00 60 0.56 Erdei et al. (2013)
H2SO4- HMF-0.2, cerevisiae biomass batch/SSCF Temperature- (Mainly glucose-122.7
impregnate Acetic acid- KE6-12 (with feed 35oC glucose) g/L and
d Steam 2.4, Formic start after 48 xylose-26.4)
explosion acid-0.4 h and stopped
pre- at 96 h of
treatment fermentation)
Wheat Dilute Furfural-2.6 7.50 Saccharomyces 20 FPU/ g Fed- pH-5.0 52.2 g/L 149.1 g/L ( 74.00 58.8 0.86 Erdei et al. (2013)
H2SO4- HMF-0.2, cerevisiae biomass batch/SSCF Temperature- (Mainly glucose-122.7
impregnate Acetic acid- KE6-12 (with feed 35oC glucose) g/L and
d Steam 2.4, Formic start after 24 xylose-26.4)
explosion acid-0.4 h and stopped

25
Journal Pre-proof

pre- at 96 h of
treatment fermentation)

Wheat Dilute Furfural-2.0 7.50 Saccharomyces 20 FPU/g Fed- pH-5.0 105.9 g/L 86.2 g/L( 97.00 43.1 4.4 Erdei et al. (2012)
(SPWS hydro H2SO4- HMF-0.1, cerevisiae biomass batch/SHCF Temperature- glucose-55.6
+ Glucose) impregnate Acetic acid- TMB3400 (Feeding start 32oC g/L, xylose-
d Steam 2.9 at 22 h and 30.6 g/L)
explosion stopped at
pre- 120 h)
treatment
Wheat Dilute Furfural-2.0 7.00 Saccharomyces 14 FPU/g Fed- pH-5.0 109.1 g/L 47.9 g/L( 95.00 43 Erdei et al. (2012)
(SPWS hydro H2SO4- HMF-0.1, cerevisiae biomass batch/SHCF( Temperature- glucose-32.3
+ Glucose) impregnate Acetic acid- TMB3400 Feeding start 32oC g/L, xylose-
d Steam 2.9 at 1 h and 15.6 g/L)
explosion stopped after
pre- 49 h)
treatment
Wheat Dilute Furfural-2.0 18.50 Saccharomyces 14 FPU/g Batch/SHCF pH-5.0 320 g/L 121.6 g/L( 87.00 53.3 0.74 Erdei et al. (2012)
(SPWS H2SO4- HMF-0.1, cerevisiae biomass Temperature- glucose-101.7
hydrol + impregnate Acetic acid- TMB3400 32oC g/L, xylose-
Starch d Steam 2.9 19.9 g/L)
hydrol) explosion
pre-
treatment
Wheat Steam Not reported 15.00 Saccharomyces 12 FPU/g Fed-batch pH-5.0 Arvinius et al.
explosion cerevisiae biomass SSF Temperature- (2010)
pre- 32oC
treatment

26
Journal Pre-proof

4.0 Techno-economic evaluation of integrated 1G2G bioethanol production

Studies on techno-economic evaluation of integrated 1G2G bioethanol production from


sugarcane using SHF with emphasis on integration at both distillation and fermentation steps
have been extensively investigated as shown in Table 5 mostly by Dias and co-workers. The
effects of acid catalyled-steam explosion pre-treatment of the lignocellulosic biomass on the
downstream processing (hydrolysis and fermentation steps) were considered by all the authors
except Dias et al. (2009) who used three-step process (separated pre-hydrolysis, delignification
and hydrolysis) in the pre-treatment stage. This further established the popularity of using
steam explosion pre-treatment of lignocellulose biomass despite the challenges such as
generation of sugar degradation products which can be reduced by optimizing the process.

The significance of integrating 1G and 2G-SHF bioethanol production from techno-economics


point of view has been widely investigated. Different scenarios such as stand-alone 1G and 2G
processes, 1G with CHP, stand-alone 2G ethanol production from co-fermentation of xylose
and glucose sugars derived from hydrolysis of pre-treated biomass, stand-alone 2G ethanol
production from fermentation of only glucose sugars with option of biogas production from
the xylose sugars, integrated 1G2G process using all the bagasse for ethanol production,
integrated 1G2G process using part of the bagasse as fuel in boiler for electricity generation
and flexible 1G2G process producing both bioethanol and electricity were investigated.
Amongst the different scenarios considered, integrated 1G2G processes using all the sugars for
ethanol production has the best economic performance (Dias et al., 2012a and Furlan et al.,
2013). However, this scenario did not have good environmental impact based on the life cycle
assessment compared to flexible integrated 1G2G process producing both electricity and
ethanol which has the highest avoidable carbon dioxide emissions compared to other scenarios
although changes in bioethanol prices had more impact on the IRR when compared with
increase in electricity spot market prices (Dias, 2013b).

Even though integration of 1G and 2G-SHF at fermentation step has been extensively
considered by Dias and co-workers, issues such as, effects of solid loading which often lead to
glucose accumulation which consequently result to end-product inhibition was not considered.
Also, the assumptions of high glucose yield made in the techno-economics studies done by
Dias and co-workers need to be experimentally verified. The use of two separate reactors for
hydrolysis and fermentation in integrated 1G and 2G-SHF processes could significantly
increase investment cost compare to integrated 1G and 2G-SSF bioethanol production process.
It is obvious from Tables 5 that the application of SSF for solids hydrolysis and fermentation
steps in 2G bioethanol production integrated with 1G sugar stream is scanty. The experimental
data from the integration of IG and 2G at the fermentation stage can be employed in modelling
software such as Aspen Plus software to build models that will be compared with literatures
obtained from integrated 1G2G bioethanol production from sugarcane using SHF for solids
hydrolysis and fermentation.

27
Journal Pre-proof

Macrelli et al. (2012) compared integration of 1G and 2G bioethanol production process at the
fermentation and distillation steps in their techno-economic studies. The result show that the
scenario integrated at the fermentation step has the highest MESP. The authors also observed
that the combined 1G and 2G ethanol production was economically feasible for all the scenario
simulated when the average ethanol selling price was 0.53US$/L, despite lower internal return
on investment (IRR< 21.5 %). It was also observed that the addition of leaves almost doubled
the amount of ethanol in 2G ethanol. Electricity price and enzyme cost were found to have
significant effect on the 2G ethanol production cost if part of the leaves and bagasse are not
considered as fuel in boiler for combine heat and power cogeneration. The authors concluded
that integration of 1G and 2G-SHF at the fermentation steps offers better economic prospect
compared to integration of 1G and 2G-SHF at the distillation steps. The technical and economic
feasibility of co-production of ethanol, lactic acid and electricity from sugarcane bagasse has
been investigated by Mandegari et al. (2017). Scenarios such as ethanol and electricity
production; lactic acid and electricity production; ethanol from cellulose, lactic acid from
hemicellulose and electricity production; ethanol from hemicellulose, lactic acid from cellulose
and electricity production were investigated. The authors revealed that the co-production of
lactic acid and electricity was the most economically attractive scenario however with the most
environmental burden compared to other scenarios. In a situation where the economic and
environmental benefits is both desired, the biorefinery concept with the production of ethanol
from hemicellulose, lactic acid and electricity would be the best scenarios based on the
indicators used. Since bioethanol production from 2G feedstock is still at the early stage of
commercialization, the full maturity of the 2G biorefinery concept can be rapidly attained by
integrating the 2G feedstock with the well matured bioethanol production from 1G feedstock.
Hence, the maximum synergistic benefits of integrating 1G2G can be achieved.

As shown in Figure 4, economic indicators of the standalone 2G and the integrated 1G2G are
measured in terms of the IRR and MESP and it varies depending on the nature of the bioethanol
production facility. Amongst the different scenarios presented, 1G2G bioethanol production
integrated at the fermentation stage using sugarcane juice and bagasse as feedstocks presented
the highest IRR (Macrelli et al., 2012). On the other hand, the IRR standalone 2G bioethanol
production using sugarcane bagasse and trash has a lower IRR compared to the integrated
1G2G bioethanol production. The higher IRR reported could be as a result of using the whole
component of the crop (juice, bagasse and straw) for bioethanol production. This will
eventually be translated to higher ethanol productivity and yield. However, the IRR for the
integrated 1G2G bioethanol production is dependent on the type of integration point. This is
evident from the variation in the IRR reported for integrated 1G2G bioethanol at CHP and
distillation (Dias at al., 2013a; Furlan et al., 2012). In addition, the integrated 1G2G bioethanol
production has a lower MESP compared to the stand 2G bioethanol production. This can be
explained in terms of the economic benefits from the synergy obtained from integrating the
standalone 2G bioethanol production with 1G production facility.

28
Journal Pre-proof

1G2G

1G2G
1G2G

1G2G
1G2G

2G

2G
1G2G

2G
IRR (%)

2G
1G2G

Macreli et al., (2012)

Dias et al., (2012b)


Furlan et al., (2013)

Dias et al., (2012b)


Dias et al., (2013c)
Macrelli et al., (2014)

Macreli et al., (2012)

Dias et al., (2012b)


Dias et al., (2012)

Sassner et al., (2008)


Dias et al., (2011a)

Figure 4: Economic performances of the simulated 2G and 1G2G in terms of IRR and MESP

The efficiency, flexibility, profitability and sustainability of the 1G2G bioethanol production
from sugar-based feedstock can be enhanced using Pinch technology as reported by Martinez-
Hernandez et al. (2013). According to Ponce-Ortega et al. (2008), pinch analysis is a process
intensification methodology that involves series of analysis performed to optimize the
utilization of energy and material utilities. Integration process by pinch analysis provides
opportunity for resourceful combination of the different pathways in the process stream for
efficient utilization of materials and energy. Martinez-Hernandez et al. (2013), employed mass
pinch analysis for a biorefinery producing bioethanol from wheat with co-production of
arabinoxylan. The bioethanol produced from the wheat was partly utilized for precipitation of
the arabinolans extract, washing and purification of wheat bran. Based on the pinch analysis,
the utilization of the bioethanol product was reduced by 94 % thereby leading to prevention of
significant loss of revenue. Besides the efficient utilization of materials using integration
process, energy recovery can also be maximized as reported by Ensinas et al. (2007). Based on
the work of Ensinas et al. (2007), thermal process integration approach was employed to
enhance energy recovery in sugar, ethanol and electricity production process thereby creating
room for an increased electricity production. Based on the thermal integration analysis, the
process steam requirement for the sugar and bioethanol process was significantly reduced
29
Journal Pre-proof

thereby facilitating a more efficient use of energy. The authors concluded that electricity
production using sugarcane bagasse and trash can be increased significantly by the reduction
of the process steam. In a similar study, Modarresi et al. (2012) employed pinch analysis for
thermal integration of combined bioethanol, biogas, heat and power production from wheat
straw as illustrated in Figure 5. The illustrated process flow involved the integration of three
different sections namely bioethanol production, biogas production and combined heat and
power production. Based on the pinch analysis of the process, the hot and cold utility was
reduced by 40 % via the synthesized heat exchanger network. Although, extensive studies have
shown that thermal integration using Pinch analysis has the potential to decrease hot and cold
utilities demand thereby boosting energy consumption during bioethanol production, however,
these studies were limited to integrated concept using 2G feedstock such as sugarcane bagasse,
corn stover and wheat straw. To maximize the full benefits of thermal integration using Pinch
technology in bioethanol production, the 1G2G bioethanol production must be given full
attention.

Figure 5: Combined production of ethanol, biomethane, heat and power from straw
(Reproduced with permission from Modarresi et al. (2012), Copyright 2012 Elsevier)

30
Journal Pre-proof
Table 5: Summary of literature on techno-economic study for 2G and integrated 1G2G bioethanol production

Feedstock Feedsto Process Investigat Pre- Hydrolysi Enzy Feedi Solid Extent Ethanol Ethanol Yield Yield Productivity Overall Economi Surplus Referen
used ck scale type ion treatment s- me ng loadin of Producti concentrat of of of the energy c electrici ce
method method Fermentat dosag strate gs (%) sugar on ion ethanol ethanol process efficien performa ty
ion e gy utilizati on on (EtOH/liter.h cy of nce (IRR, (kWh/T
method on (%) sugar sugar our) the NPV, C)
fed into consum process MESP)
the ed in (%)
fermen the
ter (%) ferment
er (%)
Sugarcane 3960 1G2G Modelling Steam SHF n.r Fed- 20 95 585 L/ 123 g/L 95 100 15 g/L.h 64.70 MESP- 56 Macrell
TC/day with explosion batch dry ton 0.49 i et al.
Aspen pre- SC US$/L (2014)
Plus treatment IRR-15.6
%
Sugarcane Not 2G Modelling Steam SHF n.r Fed- 20 95 311 n.r 95 100 n.r n.r n.r n.r Macrell
bagasse + reporte with explosion batch L/dry i et al.
trash d Aspen pre- ton SC (2014)
Plus treatment
Sugarcane 12000 1G2G Modelling Steam SHF 15 Fed- 15 90 106.7 n.r 90 100 n.r 62 IRR-13.7 80 Dias et
TC/day with explosion FPU/ batch L/dry % al.
Aspen + g ton SC (2013c)
Plus delignificat
ion
Sugarcane 2460 2G Modelling Steam SHF 15 Fed- 15 90 24.4 n.r 82 91 n.r n.r n.r n.r Dias et
bagasse + ton with explosion FPU/ batch L/dry al.
trash /day Aspen + g ton SC (2013c)
Plus delignificat
ion
Sugarcane 12000 1G2G Modelling Dilute SHF 20 Not Not 90 91.6 L/ 240 g/L 88 100 26.7 g/L.h IRR-8.3 Furlan
TC/day with H2SO4 pre- FPU/ report report dry ton % et al.
EMSO treatment g ed ed SC (2013)
software
Whole wheat Not 1G2G Experime H2SO4 - SSF 15 Batch Not n.r 59.5 g/L 95 100 0.98 g/L.h n.r n.r n.r Erdei et
reporte ntation catalyzed FPU/ report al.
d steam pre- g er (2013)
treatment
Sugarcane 12000 1G2G Modelling Steam SHF 15 Fed- 15 n.r 115.7L/ n.r 90 100 n.r 73.8 IRR-18.1 91.7 Dias et
ton with explosion FPU/ batch TC % al.
/day Aspen pre- g (2013b)
Plus treatment
followed
by alkaline
delignificat
ion
Sugarcane 3936 1G2G Modelling H3PO4 SHF 15 n.r 14 n.r 148 n.r 95 100 n.r 66.30 MESP- 106 Macrell
ton/day using catalyzed FPU/ L/ton- 0.40 i et al.
Aspen steam g DM dSC US$/L (2012)

31
Journal Pre-proof

Plus explosion and IRR-18.8


24 IU/ %
Gg
DM
Bagasse + Not 2G Modelling H3PO4 SHF NR n.r 14 n.r 102 n.r n.r n.r n.r n.r MESP- Macrell
trash reporte using catalyzed L/ton- 0.78 i et al.
d Aspen steam dSC US$/L (2012)
Plus explosion
Sugarcane 12000 1G2G Modelling Steam SHF n.r Fed- 15 n.r 116 n.r 90 100 n.r 66.70% IRR-16.8 81 Dias et
ton/day using explosion batch L/TC % al.
Aspen with MESP - (2012a)
Plus delignificat 0.33$/L
ion
Bagasse + 1770 2G Modelling Steam SHF n.r NR 15 n.r 35 L/TC n.r 82 n.r n.r n.r n.r 42 Dias et
trash kg/day, using explosion al.
dry Aspen (2012a)
basis Plus
Sugarcane 12000 1G2G Modelling Steam SHF n.r Fed- 15 n.r 113.7 n.r 90 n.r n.r n.r n.r 116 Dias et
ton/day using explosion batch L/TC al.
Aspen with (2012b)
Plus delignificat
ion
Bagasse + 1427 2G Modelling Steam SHF n.r n.r 15 n.r 31.9 n.r n.r n.r n.r 12.60 IRR-10 42 Dias et
leave ton/day using explosion L/TC % al.
, dry Aspen with or (2012b)
basis Plus without
delignificat
ion
Sugarcane 12000 1G2G Modeling Steam SHF 15 Fed- 15 92% 131 n.r 92 100 n.r 58.90% 1RR-18.4 72.7 Dias et
ton using explosion FPU/ batch L/TC % al.
/day SuperPro pre- g DM MESP- (2011a)
Designer treatment and 0.315
24 IU/ US$ /L
Gg
DM
Sugarcane 2460 2G Modelling Steam SHF 15 Fed- 15 92 41.7 48.96 g/L 92 100 2.04 g/L.h n.r IRR-15.9 none Dias et
bagasse + ton using explosion FPU/ batch L./TC % al.
trash /day SuperPro pre- g DM (2011a)
Designer treatment and
24 IU/
g DM
Corn Stover 200000 2G Experime SO2 SSF 15 n.r 10 306 n.r 78 n.r n.r 52 n.r n.r Sassner
of dry ntation + catalysed FPU/ L/ton of et al.
tons of Modelling steam g raw (2008)
materia using explosion material
ls per Aspen
annum Plus
Salix(Hardw 200000 2G Experime SO2 SSF 15 n.r 9 314 n.r 73 n.r n.r 56 n.r n.r Sassner
ood) of dry ntation + catalysed FPU/ L/ton of et al.
tons of Modelling steam g raw (2008)

32
Journal Pre-proof

materia using explosion material


ls per Aspen
annum Plus
Pruce(Softwo 200000 2G Experime SO2 SSF 15 n.r 10 315 40 g/L 69 90 0.56 g/L.h 52 MESP- n.r Sassner
od) of dry ntation + catalysed FPU/ L/ton of 0.65 US$ et al.
tons of Modelling steam g raw /L (2008)
materia using explosion material
ls per Aspen
annum Plus
Softwood 12000 2G Experime SO2 SSF 15 Fed- 8.4 0.236 35 g/L 71 n.r 0.486 g/L.h 54 n.r n.r Wingre
kg ntation + catalysed FPU/ batch L/kg of n et al.
DM/h Modelling steam g raw (2008)
using explosion material
Aspen
Plus
n.r: not reported

Note: n.r: not reported

33
Journal Pre-proof
5.0 Environmental impact of the integrated 1G2G bioethanol production

Beside the techno-economic consideration, it is worthwhile to also have an overview of the


environmental impacts of the integrated 1G2G bioethanol production. Bioethanol production
using 1G feedstocks has been reported to contribute to large extent, greenhouse gases emission
which invariably have significant environmental impact. The major environmental impact has
been reported to emanate from land use change and water utilization during cultivation of the
crops (Borrion et al., 2012) Although, extensive review of the environmental impact of 2G
bioethanol production revealed that, comparatively 1G bioethanol production has a higher
environmental impact compared to that of 2G (Gerbrandt et al., 2016). However, integrating
1G and 2G production facilities could create a significant synergy that could climax in lower
environmental impact.

Junqueira et al. (2012) revealed that the integration of 2G bioethanol production from
lignocellulosic materials with 1G -production facilities promote environmental sustainability.
Amongst the different scenarios considered by the authors, the integrated 1G2G was adjudged
to have the lowest environmental impacts. Some of the environmental impact overcome by
1G2G integration include abiotic depletion, acidification, human toxicity, freshwater aquatic
ecotoxicity, terrestrial ecotoxicity. The integrated 1G2G with steam explosion pre-treatment
resulted in the lowest environmental impact compared to using other pre-treatment methods
such as hydrothermal, acid pre-treatment and alkaline pre-treatment. The delignification of the
lignocellulose using alkaline was found to have a high environmental impact compare to other
processes. This could be as a result of huge amount of alkaline require for the delignification
process. The utilization of pentose for ethanol production rather than bio-digesting it to produce
biogas could also reduce environmental impact of 1G2G integrated process and enhance the
technical and economic viability since all the sugars obtained from the hydrolysis of the
cellulose and hemicellulose will be utilized. However, one major challenge is obtaining suitable
microorganism for the pentose fermentation. In a related study, Falano et al. (2014) reported
environmental sustainability of bioethanol production from integration point of view. The
author concluded that bioethanol produced from 1G2G integrated facilities has a better
environmental sustainability compared to fossil fuel, 1G bioethanol and 2G bioethanol. The
integrated 1G2G was observed to save up to 87 % of greenhouse emissions compared to petrol
per MJ of fuel.

Gnansounou et al. (2015) investigated the environmental impact of four different scenarios of
1G2G bioethanol production. The scenarios include 1G2G integration involving co-
fermentation of hexose and pentose sugar, 1G2G integration involving CHP, 1G2G integration
involving fermentation of only hexose sugar while pentose sugar is utilized for syrup used in
animal feed. The study revealed that IG2G integration involving the utilization of all the sugars
as well inclusion of residues for bioethanol production have the lowest environmental impact.
However, the economic analysis revealed that the IG2G integration scenario with the least
environmental impact proved to be the most expensive. Hence, further research must focus on
the trade-off between technical, economic and environmental viability of integrated 1G2G
bioethanol production.

34
Journal Pre-proof

6.0 Concluding remarks: opportunities and potential pathways for further study

Bioethanol production from 1G feedstock is a proven technology which have been extensively
investigated. The process has the advantages of high ethanol productivity and yield. There is
however issues of food-to-fuel debate and high environmental impacts. Bioethanol production
from 2G feedstocks have enjoyed tremendous attention due to the abundance of lignocellulosic
biomass resources and its prospects as a means of producing cleaner and environmentally
friendly biofuel. However, the development of the 2G process is still faced with a lot of
challenges mainly from the pre-treatment stage.

Integrating 2G bioethanol production with 1G in a single facility provides technological,


economic and environmental benefits compared to the stand-alone 2G bioethanol production
process (Mupondwa et al., 2018). Since one of the drawback in 2G bioethanol production is
high initial investment cost together with high cost of the end-product, integrating with 1G
bioethanol production might leads to reduction in capital and operating cost, as well as cost of
the end-product of the integrated process compared to stand-alone 2G process (Dias et al.,
2012). Also, 1G bioethanol production process is known to have reached advanced
technological stage in terms of infrastructure and processing while 2G bioethanol production
processes is still at the early stage of development. Integrating these processes might reduce
high risks and uncertainties in investment, thus providing opportunity for commercialization of
a viable bioethanol production process.

The high cost of feedstock which accounted for about 30 % of production cost in 1G bioethanol
production process, as well as high cost of logistics (e.g. transportation of bagasse and
agricultural residues) associated with 2G bioethanol production process can be reduced through
integration since the whole agricultural crop will be used as feedstock in the integration process.
In this case juice, bagasse, and straw thereby enabling the sharing of the cost of feedstock,
production and logistics. This will help in reducing the investment cost and increase the yield
and productivity of bioethanol since operations such as sugar concentration, fermentation,
distillation, and utilities will be integrated. The integration will also lead to significant reduction
in energy demand and capital cost per ton fuel produced in 2G (Macrrelli et al., 2012, Dias et
al., 2013a). The difficulty to attain high concentrations of ethanol in 2G process as a result of
high concentrations of sugar degradation compounds (inhibitors) caused by high severity of the
biomass pre-treatment can be overcome in the integrated process (van Zyl et al., 2011). For
example, the sugar syrup stream (1G) can be used to dilute the cellulose-derived feed (2G),
whereby the concentration of inhibitors in the 2G stream is reduce while simultaneously
increasing the final ethanol concentration in the 1G2G fermentation product. The required
energy for distillation would be reduced in the integrated process thereby reducing the overall
cost of production.

Integrated 1G2G bioethanol production from sugar-based crops present higher rates of
bioethanol production, attractive economic and more environmentally friendly when compare
to a stand-alone 2G bioethanol plant. Integrated 1G2G bioethanol production from whole wheat
using SSF process offers better performance in terms of ethanol yield, ethanol concentrations
and productivity compared to integrated 1G2G bioethanol production using SHF. The effective
35
Journal Pre-proof
application of SSF for bioethanol conversion from cellulose in the integrated process will
reduce number of process stages, capital and operating costs. Further research needs to be
conducted on the comparison of the techno-economic assessment of integrated 1G2G at the
SHF and SSF in order to justify assertion that 1G2G integration at SSF might be economical
viable than the integration at SHF. It will also enable the constraints of glucose accumulation
which often leads to low ethanol yield in SHF protocol to be overcome. Higher ethanol yield
was obtained for integrated 1G and 2G-SSF bioethanol production process compared to
integrated 1G and 2G-SHF bioethanol production process.

Integration of 1G and 2G sugar streams and the adoption of the SSCF strategy will enable full
utilization of the fermentable sugars which will result into higher ethanol concentration and
ethanol yield thereby reducing the energy demands during distillation. The integration of 1G2G
will provide further opportunity for the fermentation process development and optimization
through effective recycling of yeast (cell concentration), variation in substrate loading and
variation of mixing ratios of the sugar streams. In the light of this review, there are opportunity
for the development of strategies to integrate 1G and 2G bioethanol production from sugarcane,
sweet sorghum, and wheat using the SSF and SSCF strategies for solid hydrolysis of pre-treated
lignocellulose biomass and sugar fermentation to ethanol. The integration at the fermentation
step enables optimum biological benefits associated with this step. A further study should focus
on the incorporation of both biochemical and thermochemical route of bioethanol production
into the integrated 1G2G facilities in order to maximize the benefits of both processes. This
would enable the tackling of the series of technological constraints that could arise from the
integration processes and eventually pave way for possibilities of commercial reality of the
integrated 1G2G bioethanol production in the future.

Acknowledgement

The authors acknowledge the support of National Energy University, Malaysia through
BOLD2025 researchers grant (10436494/B2019141)

References

Akinci, B., Kassebaum, P.G., Fitch, J. V., Thompson, R.W., 2008. The role of bio-fuels in
satisfying US transportation fuel demands. Energy Policy 36, 3485–3491.
doi:10.1016/j.enpol.2008.05.021
Alfani, F., Gallifuoco, A., Saporosi, A., Spera, A., Cantarella, M., 2000. Comparison of
SHF and SSF processes for the bioconversion of steam-exploded wheat straw. J. Ind.
Microbiol. Biotechnol. 25, 184–192. doi:10.1038/sj.jim.7000054
Amores, I., Ballesteros, I., 2013. Ethanol Production from Sugarcane Bagasse Pretreated by
Steam Explosion. Electron. J. Energy Environ. 1, 25–36. doi:10.7770/ejee-V1N1-art486.
Appiah-Nkansah, N.B., Zhang, K., Rooney, W., Wang, D., 2018. Ethanol production from
mixtures of sweet sorghum juice and sorghum starch using very high gravity
fermentation with urea supplementation. Ind. Crops Prod. 111, 247–253.
doi:10.1016/j.indcrop.2017.10.028
Arora, R., Sharma, N.K., Kumar, S., Sani, R.K., 2019. Lignocellulosic Ethanol: Feedstocks
and Bioprocessing, in: Ray, C.R., Ramachandra, S. (Ed.) Bioethanol Production from

36
Journal Pre-proof
Food Crops: Sustainable Sources, Intervensions, and Challenges. Acedemic Press,
Oxford, United Kindom 165–185. doi:10.1016/b978-0-12-813766-6.00009-6
Arvinius, Emelie., Bondesson, Pia-Maria., Komorin, Ilia., Navasa, Maria., and Svensson, P.,
2010. Integration of first and second generation of ethanol production from wheat.
Working paper, Lund University, Sweeden. 1-61
Ask, M., Olofsson, K., Di Felice, T., Ruohonen, L., Penttilä, M., Lidén, G., Olsson, L., 2012.
Challenges in enzymatic hydrolysis and fermentation of pretreated Arundo donax
revealed by a comparison between SHF and SSF. Process Biochem. 47, 1452–1459.
doi:10.1016/j.procbio.2012.05.016
Badiei, M., Asim, N., Jahim, J.M., Sopian, K., 2014. Comparison of Chemical Pretreatment
Methods for Cellulosic Biomass. APCBEE Procedia 9, 170–174.
doi:10.1016/j.apcbee.2014.01.030
Banerjee, S., Environmental, N., 2010. Commercializing lignocellulosic bioethanol :
technology bottlenecks and possible remedies. Biofuel, Bioprod. Biorefining 4, 77–93.
doi:10.1002/bbb.188
Baral, N.R., Shah, A., 2017. Comparative techno-economic analysis of steam explosion,
dilute sulfuric acid, ammonia fiber explosion and biological pretreatments of corn stover.
Bioresour. Technol. 232, 331–343. doi:10.1016/j.biortech.2017.02.068
Bechara, R., Gomez, A., Saint-Antonin, V., Schweitzer, J.M., Maréchal, F., Ensinas, A.,
2018. Review of design works for the conversion of sugarcane to first and second-
generation ethanol and electricity. Renew. Sustain. Energy Rev. 91, 152–164.
doi:10.1016/j.rser.2018.02.020
Bellido, C., Bolado, S., Coca, M., Lucas, S., González-Benito, G., García-Cubero, M.T.,
2011. Effect of inhibitors formed during wheat straw pretreatment on ethanol
fermentation by Pichia stipitis. Bioresour. Technol. 102, 10868–74.
doi:10.1016/j.biortech.2011.08.128
Borrion, A.L., McManus, M.C., Hammond, G.P., 2012. Environmental life cycle assessment
of lignocellulosic conversion to ethanol: A review. Renew. Sustain. Energy Rev. 16,
4638–4650. doi.org/10.1016/j.rser.2012.04.016
Bothast, R.J., Schlicher, M. A, 2005. Biotechnological processes for conversion of corn into
ethanol. Appl. Microbiol. Biotechnol. 67, 19–25. doi:10.1007/s00253-004-1819-8
Cardona, C. A, Sánchez, O.J., 2007. Fuel ethanol production: process design trends and
integration opportunities. Bioresour. Technol. 98, 2415–57.
doi:10.1016/j.biortech.2007.01.002
Carrasco, C., Baudel, H.M., Sendelius, J., Modig, T., Roslander, C., Galbe, M., Hahn-
Hägerdal, B., Zacchi, G., Lidén, G., 2010. SO2-catalyzed steam pretreatment and
fermentation of enzymatically hydrolyzed sugarcane bagasse. Enzyme Microb. Technol.
46, 64–73. doi:10.1016/j.enzmictec.2009.10.016
Castro, E., Nieves, I.U., Mullinnix, M.T., Sagues, W.J., Hoffman, R.W., Fernández-Sandoval,
M.T., Tian, Z., Rockwood, D.L., Tamang, B., Ingram, L.O., 2014. Optimization of
dilute-phosphoric-acid steam pretreatment of Eucalyptus benthamii for biofuel
production. Appl. Energy 125, 76–83. doi:10.1016/j.apenergy.2014.03.047

37
Journal Pre-proof
Chandel, Anuj K, Lima, D.R., Mariano, A.P., 2018. Advances in Sugarcane Biorefinery
Technologies , Commercialization , Policy Issues and Paradigm Shift for Bioethanol and
By-Products, in: Chandel, Anuj Kumar, Silveira, M.H.L. (Eds.), Advances in Sugarcane
Biorefinery. Elsevier Inc. Oxford, United Kingdom. 197–212. doi:10.1016/C2015-0-
02033-0
Chaturvedi, V., Verma, P., 2013. An overview of key pretreatment processes employed for
bioconversion of lignocellulosic biomass into biofuels and value added products. 3
Biotech 3, 415–431. doi:10.1007/s13205-013-0167-8
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. doi:10.1016/j.fuproc.2016.12.007
Conde-Mejía, C. Jimenez-Gutierrez, A. & El-Halwagi, M., 2013. Assessment of
Combinations between Pretreatment and Conversion Configurations for Bioethanol
Production. ACS Sustainable Chem Eng. 1, 956–955.
Conde-Mejía, C., Jiménez-Gutiérrez, A., El-Halwagi, M., 2012. A comparison of
pretreatment methods for bioethanol production from lignocellulosic materials. Process
Saf. Environ. Prot. 90, 189–202. doi:10.1016/j.psep.2011.08.004
Cotana, F., Cavalaglio, G., Gelosia, M., Coccia, V., Petrozzi, A., Ingles, D., Pompili, E.,
2015. A comparison between SHF and SSSF processes from cardoon for ethanol
production. Ind. Crops Prod. 69, 424–432. doi.org/10.1016/j.indcrop.2015.02.064
Crago, C.L., Khanna, M., Barton, J., Giuliani, E., Amaral, W., 2010. Competitiveness of
Brazilian sugarcane ethanol compared to US corn ethanol. Energy Policy 38, 7404–7415.
doi:10.1016/j.enpol.2010.08.016
Dahnum, D., Tasum, S.O., Triwahyuni, E., Nurdin, M., Abimanyu, H., 2015. Comparison of
SHF and SSF processes using enzyme and dry yeast for optimization of bioethanol
production from empty fruit bunch. Energy Procedia 68, 107–116.
doi:10.1016/j.egypro.2015.03.238
Damartzis, T., Zabaniotou, A., 2011. Thermochemical conversion of biomass to second
generation biofuels through integrated process design—A review. Renew. Sustain.
Energy Rev. 15, 366–378. doi:10.1016/j.rser.2010.08.003
Damay, J., Boboescu, I., Duret, X., Lalonde, O., Lavoie, J., 2018. A novel hybrid fi rst and
second generation hemicellulosic bioethanol production process through steam treatment
of dried sorghum biomass. Bioresour. Technol. 263, 103–111.
doi:10.1016/j.biortech.2018.04.045
de Carvalho, A.L., Antunes, C.H., Freire, F., 2016. Economic-energy-environment analysis of
prospective sugarcane bioethanol production in Brazil. Appl. Energy 181, 514–526.
doi:10.1016/j.apenergy.2016.07.122
de Souza, C.J. a, Costa, D. a, Rodrigues, M.Q.R.B., dos Santos, A.F., Lopes, M.R., Abrantes,
A.B.P., dos Santos Costa, P., Silveira, W.B., Passos, F.M.L., Fietto, L.G., 2012. The
influence of presaccharification, fermentation temperature and yeast strain on ethanol
production from sugarcane bagasse. Bioresour. Technol. 109, 63–9.
doi:10.1016/j.biortech.2012.01.024
Deenanath, E.D., Rumbold, K., Iyuke, S., 2013. The Production of Bioethanol from Cashew
38
Journal Pre-proof
Apple Juice by Batch Fermentation Using Saccharomyces cerevisiae Y2084 and Vin13 .
ISRN Renew. Energy 2013, 1–11. doi:10.1155/2013/107851
Deesuth, O., Laopaiboon, P., Klanrit, P., Laopaiboon, L., 2015. Improvement of ethanol
production from sweet sorghum juice under high gravity and very high gravity
conditions: Effects of nutrient supplementation and aeration. Ind. Crops Prod. 74, 95–
102. doi:10.1016/j.indcrop.2015.04.068
Dehghan-Niri, R., Walmsley, J.C., Holmen, A., Midgley, P. a., Rytter, E., Dam, A.H.,
Hungria, A.B., Hernandez-Garrido, J.C., Chen, D., 2012. Nanoconfinement of Ni
clusters towards a high sintering resistance of steam methane reforming catalysts. Catal.
Sci. Technol. 2, 2476. doi:10.1039/c2cy20325a
Della-Bianca, B. E., Basso, T. O, Stambuk, U. B., Basso, C.L., Gombert, A.K., 2013. What
do you know about the yeast strains from Brazilian fuel ethanol industry? Appl.
Microbiol. Biotechnol. 97, 979–991. doi:10.1007/s00253-012-4631-x
Demirbas, A, 2007. Progress and recent trends in biofuels. Prog. Energy Combust. Sci. 33,
1–18. doi:10.1016/j.pecs.2006.06.001
Dessie, W., Xin, F., Zhang, W., Zhou, J., Wu, H., Ma, J., Jiang, M., 2019. Inhibitory effects
of lignocellulose pretreatment degradation products (hydroxymethylfurfural and furfural)
on succinic acid producing Actinobacillus succinogenes. Biochem. Eng. J. 150, 107263.
doi:10.1016/j.bej.2019.107263
Di Donato, P., Finore, I., Poli, A., Nicolaus, B., Lama, L., 2019. The production of second
generation bioethanol: The biotechnology potential of thermophilic bacteria. J. Clean.
Prod. 233, 1410–1417. doi:10.1016/j.jclepro.2019.06.152
Dias, M.O.S., Ensinas, A. V., Nebra, S. A., Maciel Filho, R., Rossell, C.E.V., Maciel,
M.R.W., 2009. Production of bioethanol and other bio-based materials from sugarcane
bagasse: Integration to conventional bioethanol production process. Chem. Eng. Res.
Des. 87, 1206–1216. doi:10.1016/j.cherd.2009.06.020
Dias, M.O.S., Junqueira, T.L., Cavalett, O., Cunha, M.P., Jesus, C.D.F., Mantelatto, P.E.,
Rossell, C.E.V., Maciel Filho, R., Bonomi, A., 2013. Cogeneration in integrated first and
second generation ethanol from sugarcane. Chem. Eng. Res. Des. 91, 1411–1417.
doi:10.1016/j.cherd.2013.05.009
Dias, M.O.S., Junqueira, T.L., Jesus, C.D.F., Rossell, C.E.V., Maciel Filho, R., Bonomi, A.,
2012. Improving second generation ethanol production through optimization of first
generation production process from sugarcane. Energy 43, 246–252.
doi:10.1016/j.energy.2012.04.034
Ding, S., Tan, T., 2006. l-lactic acid production by Lactobacillus casei fermentation using
different fed-batch feeding strategies. Process Biochem. 41, 1451–1454.
doi:10.1016/j.procbio.2006.01.014
Duque, A., Manzanares, P., Ballesteros, I., Ballesteros, M., 2016. Chapter 15 - Steam
Explosion as Lignocellulosic Biomass Pretreatment, in: Mussatto, S.I. (Ed.), Biomass
Fractionation Technologies for a Lignocellulosic Feedstock Based Biorefinery. Elsevier,
Amsterdam, Netherlands 349–368. doi.org/10.1016/B978-0-12-802323-5.00015-3
Dutta, K., Daverey, A., Lin, J., 2014. Evolution retrospective for alternative fuels : First to
fourth generation. Renew. Energy 69, 114–122. doi:10.1016/j.renene.2014.02.044
39
Journal Pre-proof
Europian Biofuel Technology Platform, 2013. Cellulosic Ethanol (CE). EBTP Newsletter.
www.biofuelstp.eu/cell_ethanol.html#brazil (accessed 20 August 2018)
EIA, 2017. Global Transportation Energy Consumption: Examination of Scenarios to 2040
using ITEDD. US Department of Energy, Washington. 1-35
Ensinas, A. V., Nebra, S.A, Lozano, M.A., Serra, L.M., 2007. Analysis of process steam
demand reduction and electricity generation in sugar and ethanol production from
sugarcane. Energy Convers. Manag. 48, 2978–2987.
doi:10.1016/j.enconman.2007.06.038
Erdei, B., Barta, Z., Sipos, B., Réczey, K., Galbe, M., Zacchi, G., 2010. Ethanol production
from mixtures of wheat straw and wheat meal. Biotechnol. Biofuels 3, 16.
doi:10.1186/1754-6834-3-16
Erdei, B., Hancz, D., Galbe, M., Zacchi, G., 2013. SSF of steam-pretreated wheat straw with
the addition of saccharified or fermented wheat meal in integrated bioethanol production.
Biotechnol. Biofuels 6, 169. doi:10.1186/1754-6834-6-169
Falano, T., Jeswani, H.K., Azapagic, A., 2014. Assessing the environmental sustainability of
ethanol from integrated biorefineries. Biotechnol. J. 9, 753–765.
doi:10.1002/biot.201300246
Fan, Y. Van, Perry, S., Klemes, J.J., Lee, C.T., 2018. A review on air emissions assessment :
Transportation. J. Clean. Prod. 194, 673–684. doi:10.1016/j.jclepro.2018.05.151
Ferreira, J.A., Brancoli, P., Agnihotri, S., Bolton, K., Taherzadeh, M.J., 2018. A review of
integration strategies of lignocelluloses and other wastes in 1st generation bioethanol
processes. Process Biochem. 75, 173–186. doi:10.1016/j.procbio.2018.09.006
Filip, O., Janda, K., Kristoufek, L., Zilberman, D., 2017. Food versus fuel: An updated and
expanded evidence. Institute of Economic Studies, Charles University Prague Working
Paper 1-40 doi.org/10.1016/j.eneco.2017.10.033
Finkbeiner, M., 2014. Indirect land use change – Help beyond the hype? Biomass and
Bioenergy 62, 218–221. doi:10.1016/j.biombioe.2014.01.024
Fish, W.W., Bruton, B.D., Russo, V.M., 2009. Watermelon juice: a promising feedstock
supplement, diluent, and nitrogen supplement for ethanol biofuel production. Biotechnol.
Biofuels 2, 1-9 doi:10.1186/1754-6834-2-18
Friedl, A., 2019. Bioethanol from Sugar and Starch, in: Kaltschmitt, M. (Ed.), Energy from
Organic Materials (Biomass): A Volume in the Encyclopedia of Sustainability Science
and Technology, Second Edition. Springer New York, US. 905–924. doi:10.1007/978-1-
4939-7813-7_432
Gavahian, M., Munekata, P.E.S., Eş, I., Lorenzo, J.M., Khaneghah, A.M., Barba, F.J., 2018.
Emerging techniques in bioethanol production : from distillation to waste valorization.
Green Chem. 21, 1171–1185. doi:10.1039/c8gc02698j
Gawande, S.B., Patil, I.D., 2018. Experimental Investigation and Optimization for Production
of Bioethanol from Damaged Corn Grains. Mater. Today Proc. 5, 1509–1517.
doi:10.1016/j.matpr.2017.11.240
Gerbrandt, K., Chu, P.L., Simmonds, A., Mullins, K.A., MacLean, H.L., Griffin, W.M.,
Saville, B.A., 2016. Life cycle assessment of lignocellulosic ethanol: a review of key
40
Journal Pre-proof
factors and methods affecting calculated GHG emissions and energy use. Curr. Opin.
Biotechnol. 38, 63–70. doi.org/10.1016/j.copbio.2015.12.021
Gibreel, A., Sandercock, J.R., Lan, J., Goonewardene, L.A., Zijlstra, R.T., Curtis, J.M.,
Bressler, D.C., 2009. Fermentation of Barley by Using Saccharomyces cerevisiae:
Examination of Barley as a Feedstock for Bioethanol Production and Value-Added
Products. Appl. Environ. Microbiol. 75, 1363–1372. doi:10.1128/AEM.01512-08
Gnansounou, E., Vaskan, P., Pachón, E.R., 2015. Comparative techno-economic assessment
and LCA of selected integrated sugarcane-based biorefineries. Bioresour. Technol. 196,
364–375. doi.org/10.1016/j.biortech.2015.07.072
Goh, C.S., Tan, K.T., Lee, K.T., Bhatia, S., 2010. Bio-ethanol from lignocellulose: Status,
perspectives and challenges in Malaysia. Bioresour. Technol. 101, 4834–4841.
doi:10.1016/j.biortech.2009.08.080
Goldemberg, J., 2007. Ethanol for a sustainable energy future. Science 315, 808–10.
doi:10.1126/science.1137013
Gomez-Flores, R., Thiruvengadathan, T.N., Nicol, R., Gilroyed, B., Morrison, M., Reid,
L.M., Margaritis, A., 2018. Bioethanol and biobutanol production from sugarcorn juice.
Biomass and Bioenergy 108, 455–463. doi:10.1016/j.biombioe.2017.10.038
Hemansi, Gupta, R., Yadav, G., Kumar, G., Yadav, A., Saini, J.K., Kuhad, R.C., 2019.
Second Generation Bioethanol Production: The State of Art, in: Srivastava, N.,
Srivastava, M., Mishra, P.K., Upadhyay, S.N., Ramteke, P.W., Gupta, V.K. (Eds.),
Sustainable Approaches for Biofuels Production Technologies: From Current Status to
Practical Implementation. Springer International Publishing, New York, US. 121–146.
doi:10.1007/978-3-319-94797-6_8
HLPE, 2013. Biofuels and food security.y. A report by the High Level Panel of Experts on
Food Security and Nutrition of the Committee on World Food Security, Rome 1-130.
Jiang, Y., Lv, Y., Wu, R., Sui, Y., Chen, C., Xin, F., Zhou, J., Dong, W., Jiang, M., 2019.
Current status and perspectives on biobutanol production using lignocellulosic
feedstocks. Bioresour. Technol. Reports 7, 100245. doi:10.1016/j.biteb.2019.100245
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.
doi:10.1016/j.biortech.2015.10.009
Junqueira, T.L., Dias, M.O.S., Cavalett, O., Jesus, C.D.F., Cunha, M.P., Rossell, C.E. V,
Filho, R.M., Bonomi, A., 2012. Economic and environmental assessment of integrated
1st and 2nd generation sugarcane bioethanol production evaluating different 2nd
generation process alternatives, in: Bogle, I.D.L., Fairweather, M. (Eds.), 22nd European
Symposium on Computer Aided Process Engineering, Elsevier, Amsterdam, Netherlands
177–181. doi.org/10.1016/B978-0-444-59519-5.50036-8
Kadhum, H.J., Rajendran, K., Murthy, G.S., 2017. Effect of solids loading on ethanol
production: Experimental, economic and environmental analysis. Bioresour. Technol.
244, 108–116. doi.org/10.1016/j.biortech.2017.07.047
Kawa-Rygielska, J., Pietrzak, W., Regiec, P., Stencel, P., 2013. Utilization of concentrate
after membrane filtration of sugar beet thin juice for ethanol production. Bioresour.
Technol. 133, 134–141. doi:10.1016/j.biortech.2013.01.070
41
Journal Pre-proof
Kim, J.Y., Lee, H.W., Lee, S.M., Jae, J., Park, Y.K., 2019. Overview of the recent advances
in lignocellulose liquefaction for producing biofuels, bio-based materials and chemicals.
Bioresour. Technol. 279, 373–384. doi:10.1016/j.biortech.2019.01.055
Krishnan, C., Sousa, L.D.C., Jin, M., Chang, L., Dale, B.E., Balan, V., 2010. Alkali-based
AFEX pretreatment for the conversion of sugarcane bagasse and cane leaf residues to
ethanol. Biotechnol. Bioeng. 107, 441–50. doi:10.1002/bit.22824
Kumar, D., Murthy, G.S., 2011. Impact of pretreatment and downstream processing
technologies on economics and energy in cellulosic ethanol production. Biotechnol.
Biofuels 4, 27. doi:10.1186/1754-6834-4-27
Kumar, P., Varkolu, M., Mailaram, S., Kunamalla, A., Maity, S.K., 2019. Biorefinery
Polyutilization Systems: Production of Green Transportation Fuels From Biomass,
Polygeneration with Polystorage for Chemical and Energy Hubs. Academic Press,
Massachusetts, US 1-583 doi:10.1016/b978-0-12-813306-4.00012-4
Laopaiboon, L., Nuanpeng, S., Srinophakun, P., Klanrit, P., Laopaiboon, P., 2009. Ethanol
production from sweet sorghum juice using very high gravity technology: Effects of
carbon and nitrogen supplementations. Bioresour. Technol. 100, 4176–4182.
doi:10.1016/j.biortech.2009.03.046
Laopaiboon, L., Thanonkeo, P., Jaisil, P., Laopaiboon, P., 2007. Ethanol production from
sweet sorghum juice in batch and fed-batch fermentations by Saccharomyces cerevisiae.
World J. Microbiol. Biotechnol. 23, 1497–1501. doi:10.1007/s11274-007-9383-x
Laser, M., Schulman, D., Allen, S.G., Lichwa, J., Antal, M.J., Lynd, L.R., 2002. A
comparison of liquid hot water and steam pretreatments of sugar cane bagasse for
bioconversion to ethanol. Bioresour. Technol. 81, 33–44.
Lau, M.W., Dale, B.E., 2010. Effect of primary degradation-reaction products from Ammonia
Fiber Expansion (AFEX)-treated corn stover on the growth and fermentation of
Escherichia coli KO11. Bioresour. Technol. 101, 7849–55.
doi:10.1016/j.biortech.2010.04.076
Li, B.-Z., Balan, V., Yuan, Y.-J., 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, 1285–92. doi:10.1016/j.biortech.2009.09.044
Li, H., Mei, X., Liu, B., Li, Z., Wang, B., Ren, N., Xing, D., 2019. Insights on acetate-ethanol
fermentation by hydrogen-producing Ethanoligenens under acetic acid accumulation
based on quantitative proteomics. Environ. Int. 129, 1–9.
doi:10.1016/j.envint.2019.05.013
Li, W.C., Zhu, J.Q., Zhao, X., Qin, L., Xu, T., Zhou, X., Li, X., Li, B.Z., Yuan, Y.J., 2019.
Improving co-fermentation of glucose and xylose by adaptive evolution of engineering
xylose-fermenting Saccharomyces cerevisiae and different fermentation strategies.
Renew. Energy 139, 1176–1183. doi:10.1016/j.renene.2019.03.028
Liang, L., Zhang, Y.P., Zhang, L., Zhu, M.J., Liang, S.Z., Huang, Y.N., 2008. Study of
sugarcane pieces as yeast supports for ethanol production from sugarcane juice and
molasses. J. Ind. Microbiol. Biotechnol. 35, 1605–1613. doi:10.1007/s10295-008-0404-z
Liu, C., Zhang, P., Zhang, S., 2014. Feeding strategies for the enhanced production of a -
arbutin in the fed-batch fermentation of Xanthomonas maltophilia BT-112. Bioprocess
42
Journal Pre-proof
Biosyst. Eng. 35, 325–329. doi:10.1007/s00449-013-0980-9
Liu, L., Zhang, Z., Wang, J., Fan, Y., Shi, W., Liu, X., Shun, Q., 2019. Simultaneous
saccharification and co-fermentation of corn stover pretreated by H2O2 oxidative
degradation for ethanol production. Energy 168, 946–952.
doi:10.1016/j.energy.2018.11.132
Liu, R., Li, J., Shen, F., 2008. Refining bioethanol from stalk juice of sweet sorghum by
immobilized yeast fermentation. Renew. Energy 33, 1130–1135.
doi.org/10.1016/j.renene.2007.05.046
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. Bioresour. Technol. 224, 307–313.
doi:10.1016/j.biortech.2016.10.075
Lozano, F.J., Lozano, R., 2018. Assessing the potential sustainability benefits of agricultural
residues: Biomass conversion to syngas for energy generation or to chemicals
production. J. Clean. Prod. 172, 4162–4169. doi:10.1016/j.jclepro.2017.01.037
Lu, J., Li, X., Yang, R., Zhao, J., Qu, Y., 2013. Tween 40 pretreatment of unwashed water-
insoluble solids of reed straw and corn stover pretreated with liquid hot water to obtain
high concentrations of bioethanol. Biotechnol. Biofuels 6, 1-11. doi:10.1186/1754-6834-
6-159
Ma, J., Shi, S., Jia, X., Xia, F., Ma, H., Gao, J., Xu, J., 2019. Advances in catalytic conversion
of lignocellulose to chemicals and liquid fuels. J. Energy Chem. 36, 74–86.
doi:10.1016/j.jechem.2019.04.026
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. Biotechnol. Biofuels 7, 26: 1-16. doi:10.1186/1754-6834-
7-26
Macrelli, S., Mogensen, J., Zacchi, G., 2012. Techno-economic evaluation of 2nd generation
bioethanol production from sugar cane bagasse and leaves integrated with the sugar-
based ethanol process. Biotechnol. Biofuels 5, 22: 1-18. doi:10.1186/1754-6834-5-22
Mączyńska, J., Krzywonos, M., Kupczyk, A., Tucki, K., Sikora, M., Pińkowska, H., Bączyk,
A., Wielewska, I., 2019. Production and use of biofuels for transport in Poland and
Brazil – The case of bioethanol. Fuel 241, 989–996. doi:10.1016/j.fuel.2018.12.116
Magnusson, H., 2006. Process simulation in Aspen Plus of an integrated ethanol and CHP
plant. M.Sc Thesis, Universitet Umea, Sweeden 1-47
Mandegari, M.A., Farzad, S., Rensburg, E. van, Gorgens, J.F., 2017. Multi-criteria analysis of
a biorefinery for co-production of lactic acid and ethanol from sugarcane lignocellulose.
Biofuels, Bioprod. Biorefining 11, 971–990. doi:10.1002/bbb
Martín, C., Klinke, H.B., Thomsen, A.B., 2007. Wet oxidation as a pretreatment method for
enhancing the enzymatic convertibility of sugarcane bagasse. Enzyme Microb. Technol.
40, 426–432. doi:10.1016/j.enzmictec.2006.07.015
Martín, C., Wu, G., Wang, Z., Stagge, S., Jönsson, L.J., 2018. Formation of microbial
inhibitors in steam-explosion pretreatment of softwood impregnated with sulfuric acid

43
Journal Pre-proof
and sulfur dioxide. Bioresour. Technol. 262, 242–250.
doi:10.1016/j.biortech.2018.04.074
Martinez-Hernandez, E., Sadhukhan, J., Campbell, G.M., 2013. Integration of bioethanol as
an in-process material in biorefineries using mass pinch analysis. Appl. Energy 104,
517–526. doi:10.1016/j.apenergy.2012.11.054
Marzo, C., Díaz, A.B., Caro, I., Blandino, A., 2019. Status and Perspectives in Bioethanol
Production From Sugar Beet, in: Ray, R.C., Ramachandran, S. (Eds.), Bioethanol
Production from Food Crops. Academic Press, Massachusetts, US 61–79.
doi.org/10.1016/B978-0-12-813766-6.00004-7
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). Bioresour. Technol. 272, 1–9. doi:10.1016/j.biortech.2018.09.143
Mekonnen, M.M., Romanelli, T.L., Ray, C., Hoekstra, A.Y., Liska, A.J., Neale, C.M.U.,
2018. Water , Energy , and Carbon Footprints of Bioethanol from the U . S . and Brazil.
Environ. Sceince Technol. 52, 14508–14518.
Mithra, M.G., Jeeva, M.L., Sajeev, M.S., Padmaja, G., 2018. Comparison of ethanol yield
from pretreated lignocellulo-starch biomass under fed-batch SHF or SSF modes. Heliyon
4, 1-31. doi:10.1016/j.heliyon.2018.e00885
Modarresi, A., Kravanja, P., Friedl, A., 2012. Pinch and exergy analysis of lignocellulosic
ethanol, biomethane, heat and power production from straw. Appl. Therm. Eng. 43, 20–
28. doi:10.1016/j.applthermaleng.2012.01.026
Mohanty, Sujit Kumar, Swain, M.R., 2019. Bioethanol Production From Corn and Wheat:
Food, Fuel, and Future, in: Ray, R.C., Ramachandran, S. (Eds.), Bioethanol Production
from Food Crops. Academic Press Massachusetts, US 45–79. doi:10.1016/C2017-0-
00234-3
Mohanty, Sujit K, Swain, M.R., 2019. Bioethanol Production From Corn and Wheat: Food,
Fuel, and Future, in: Ray, R.C., Ramachandran, S. (Eds.), Bioethanol Production from
Food Crops. Academic Press Massachusetts, US 45–59. doi.org/10.1016/B978-0-12-
813766-6.00003-5
Mohd Azhar, S.H., Abdulla, R., Jambo, S.A., Marbawi, H., Gansau, J.A., Mohd Faik, A.A.,
Rodrigues, K.F., 2017. Yeasts in sustainable bioethanol production: A review. Biochem.
Biophys. Reports 10, 52–61. doi:10.1016/j.bbrep.2017.03.003
Mupondwa, E., Li, X., Tabil, L., 2018. Integrated bioethanol production from triticale grain
and lignocellulosic straw in Western Canada. Ind. Crop. Prod. 117, 75–87.
doi:10.1016/j.indcrop.2018.02.070
Naik, S.N., Goud, V. V., Rout, P.K., Dalai, A.K., 2010. Production of first and second
generation biofuels: A comprehensive review. Renew. Sustain. Energy Rev. 14, 578–
597. doi:10.1016/j.rser.2009.10.003
Naresh Kumar, M., Ravikumar, R., Thenmozhi, S., Ranjith Kumar, M., Kirupa Shankar, M.,
2019. Choice of Pretreatment Technology for Sustainable Production of Bioethanol from
Lignocellulosic Biomass: Bottle Necks and Recommendations. Waste and Biomass
Valorization 10, 1693–1709. doi:10.1007/s12649-017-0177-6
44
Journal Pre-proof
Nicola, G. Di, Santecchia, E., Santori, G., Polonara, F., Politecnica, U., 2011. Advances in the
Development of Bioethanol : A Review. Biofuel’s Eng. Process Technol. 611–638.
Niphadkar, S., Bagade, P., Ahmed, S., 2018. Bioethanol production : insight into past ,
present and future perspectives. Biofuels 9, 229–238.
doi:10.1080/17597269.2017.1334338
Ojeda, K., Sánchez, E., El-Halwagi, M., Kafarov, V., 2011. Exergy analysis and process
integration of bioethanol production from acid pre-treated biomass: Comparison of SHF,
SSF and SSCF pathways. Chem. Eng. J. 176–177, 195–201.
doi.org/10.1016/j.cej.2011.06.083
Oliveira, C.M., Cruz, A.J.G., Costa, C.B.B., 2016. Improving second generation bioethanol
production in sugarcane biorefineries through energy integration. Appl. Therm. Eng.
109, 819–827. doi:10.1016/j.applthermaleng.2014.11.016
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, 7.
doi:10.1186/1754-6834-1-7
Palacious-Bereche, R., Ensinas, A., Nebra, S.A., 2011. Energy consumption in Ethanol
Production by Enzymatic Hydrolysis-The Integration with the Conventional Process
Using Pinch Pnalysis. Chem. Eng. Trans. 24, 1189–1194. doi:10.3303/CET1124199
Pan, X., Gilkes, N., Kadla, J., Pye, K., 2006. Bioconversion of hybrid poplar to ethanol and
co‐products using an organosolv fractionation process: Optimization of process yields.
Biotechnol. Bioeng 94, 851-861. doi:10.1002/bit.20905
Pandey, A.K., Kumar, M., Kumari, S., Kumari, P., Yusuf, F., Jakeer, S., Naz, S., Chandna, P.,
Bhatnagar, I., Gaur, N.A., 2019. Evaluation of divergent yeast genera for fermentation-
associated stresses and identification of a robust sugarcane distillery waste isolate
Saccharomyces cerevisiae NGY10 for lignocellulosic ethanol production in SHF and
SSF. Biotechnol. Biofuels 12, 40: 1-23. doi:10.1186/s13068-019-1379-x
Patni, N., Pillai, S.G., Dwivedi, A.H., 2013. Wheat as a promising substitute of corn for
bioethanol production. Procedia Eng. 51, 355–362. doi:10.1016/j.proeng.2013.01.049
Paulino, J., Souza, D., Dias, C., Eleutherio, E.C.A., Bonatto, D., Ferreira, A., 2018.
Improvement of Brazilian bioethanol production, challenges and perspectives on the
identi fi cation and genetic modi fi cation of new strains of Saccharomyces cerevisiae
yeasts isolated during ethanol process. Fungal Biol. 122, 583–591.
doi:10.1016/j.funbio.2017.12.006
Pengilly, C., García-Aparicio, M.P., Diedericks, D., Brienzo, M., Görgens, J.F., 2015.
Enzymatic hydrolysis of steam-pretreated sweet sorghum bagasse by combinations of
cellulase and endo-xylanase. Fuel 154, 352–360. doi:10.1016/j.fuel.2015.03.072
Pescarolo, S., 2013. Case study on the first advanced industrial demonstration bioethanol
plant in the EU, and how it was financed. 5th Stakeholder plenary meeting report,
Bruxelles Italy 1-15.
Phuttaro, C., Sawatdeenarunat, C., Surendra, K.C., Boonsawang, P., Chaiprapat, S., Khanal,
S.K., 2019. Anaerobic digestion of hydrothermally-pretreated lignocellulosic biomass:
Influence of pretreatment temperatures, inhibitors and soluble organics on methane yield.
Bioresour. Technol. 284, 128–138. doi:10.1016/j.biortech.2019.03.114
45
Journal Pre-proof
Ponce-Ortega, J.M., Serna-González, M., Jiménez-Gutiérrez, A., 2008. Design and
optimization of multipass heat exchangers. Chem. Eng. Process. Process Intensif. 47,
906–913. doi:10.1016/j.cep.2007.02.004
Proskurina, S., Junginger, M., Heinimö, J., Tekinel, B., Vakkilainen, E., 2019. Global
biomass trade for energy — Part 2 : Production and trade streams of wood pellets ,
liquid. Biofuel, Bioprod. Biorefining 13, 371–387. doi:10.1002/bbb.1858
Quintero, J. a, Moncada, J., Cardona, C. a, 2013. Techno-economic analysis of bioethanol
production from lignocellulosic residues in Colombia: a process simulation approach.
Bioresour. Technol. 139, 300–7. doi:10.1016/j.biortech.2013.04.048
Rajendran, K., Drielak, E., Varma, S., Muthusamy, S., Kumar, G., 2018. Updates on the
pretreatment of lignocellulosic feedstocks for bioenergy production- a review. Biomass
Convers. Biorefinery 8, 471–483. doi:10.1007/s13399-017-0269-3
Rastogi, M., Shrivastava, S., 2017. Recent advances in second generation bioethanol
production: An insight to pretreatment, saccharification and fermentation processes.
Renew. Sustain. Energy Rev. 80, 330–340. doi:10.1016/j.rser.2017.05.225
Raud, M., Kikas, T., Sippula, O., Shurpali, N.J., 2019. Potentials and challenges in
lignocellulosic biofuel production technology. Renew. Sustain. Energy Rev. 111, 44–56.
doi:10.1016/j.rser.2019.05.020
Razmovski, R., Vučurović, V., 2012. Bioethanol production from sugar beet molasses and
thick juice using Saccharomyces cerevisiae immobilized on maize stem ground tissue.
Fuel 92, 1–8. doi:10.1016/j.fuel.2011.07.046
Renewable Fuels Association, 2019. 2019 ETHANOL INDUSTRY OUTLOOK.
Rocha, G.J.M., Gonçalves, A. R., Oliveira, B.R., Olivares, E.G., Rossell, C.E.V., 2012.
Steam explosion pretreatment reproduction and alkaline delignification reactions
performed on a pilot scale with sugarcane bagasse for bioethanol production. Ind. Crops
Prod. 35, 274–279. doi:10.1016/j.indcrop.2011.07.010
Rohowsky, B., Häßler, T., Gladis, A., Remmele, E., Schieder, D., Faulstich, M., 2013.
Feasibility of simultaneous saccharification and juice co-fermentation on hydrothermal
pretreated sweet sorghum bagasse for ethanol production. Appl. Energy 102, 211–219.
doi:10.1016/j.apenergy.2012.03.039
Rudel, T.K., 2013. Food Versus Fuel: Extractive Industries, Insecure Land Tenure, and Gaps
in World Food Production. World Dev. 51, 62–70. doi:10.1016/j.worlddev.2013.05.015
Rudolf, A., Baudel., H., Zacchi, G., Hahn-Hagerdal, B. and Liden, G., 2007. Simultaneous
saccharification and fermentation of steam‐pretreated bagasse using Saccharomyces
cerevisiae TMB3400 and Pichia stipitis CBS6054. Biotechnol.BioEng 99, 783–790.
doi:10.1002/bit.21636
Saladini, F., Patrizi, N., Pulselli, F.M., Marchettini, N., Bastianoni, S., 2016. Guidelines for
emergy evaluation of first, second and third generation biofuels. Renew. Sustain. Energy
Rev. 66, 221–227. doi:10.1016/j.rser.2016.07.073
Sasaki, K., Tsuge, Y., Kawaguchi, H., Yasukawa, M., Sasaki, D., Sazuka, T., Kamio, E.,
Ogino, C., Matsuyama, H., Kondo, A., 2017. Sucrose purification and repeated ethanol
production from sugars remaining in sweet sorghum juice subjected to a membrane

46
Journal Pre-proof
separation process. Appl. Microbiol. Biotechnol. 101, 6007–6014. doi:10.1007/s00253-
017-8316-3
Sewsynker-Sukai, Y., Gueguim Kana, E.B., 2018. Simultaneous saccharification and
bioethanol production from corn cobs: Process optimization and kinetic studies.
Bioresour. Technol. 262, 32–41. doi:10.1016/j.biortech.2018.04.056
Shi, N., Liu, D., Huang, Q., Guo, Z., Jiang, R., Wang, F., Chen, Q., Li, M., Shen, G., Wen, F.,
2019. Product-oriented decomposition of lignocellulose catalyzed by novel
polyoxometalates-ionic liquid mixture. Bioresour. Technol. 283, 174–183.
doi:10.1016/j.biortech.2019.03.048
Shuai, L., Yang, Q., Zhu, J.Y., Lu, F.C., Weimer, P.J., Ralph, J., Pan, X.J., 2010.
Comparative study of SPORL and dilute-acid pretreatments of spruce for cellulosic
ethanol production. Bioresour. Technol. 101, 3106–14.
doi:10.1016/j.biortech.2009.12.044
Slupska, M., Bushong, D., 2019. Lessons from Commercialization of Cellulosic Ethanol - A
POET Perspective. Biofuels, Bioprod. Biorefining 13, 857–859. doi:10.1002/bbb.2033
Song, Y., Cho, E.J., Park, C.S., Oh, C.H., Park, B.J., Bae, H.J., 2019. A strategy for sequential
fermentation by Saccharomyces cerevisiae and Pichia stipitis in bioethanol production
from hardwoods. Renew. Energy 139, 1281–1289. doi:10.1016/j.renene.2019.03.032
Subsamran, K., Mahakhan, P., Vichitphan, K., Vichitphan, S., Sawaengkaew, J., 2019.
Potential use of vetiver grass for cellulolytic enzyme production and bioethanol
production. Biocatal. Agric. Biotechnol. 17, 261–268. doi:10.1016/j.bcab.2018.11.023
Sudiyani, Y., Dahnum, D., Burhani, D., Putri, A.M.H., 2019. Chapter 10 - Evaluation and
comparison between simultaneous saccharification and fermentation and separated
hydrolysis and fermentation process, in: Basile, A., Dalena, F. (Eds.), Second and Third
Generation of Feedstocks. Elsevier, Amsterdam, Netherlands 273–290.
doi.org/10.1016/B978-0-12-815162-4.00010-0
Szambelan, K., Nowak, J., Szwengiel, A., Jeleń, H., Łukaszewski, G., 2018. Separate
hydrolysis and fermentation and simultaneous saccharification and fermentation methods
in bioethanol production and formation of volatile by-products from selected corn
cultivars. Ind. Crops Prod. 118, 355–361. doi:10.1016/j.indcrop.2018.03.059
Taber, A., 2013. Ethanol Production in Brazil.
www.worldfoodprize.org/documents/filelibrary/images/youth_programs/research_papers
/2007_papers/Abraham_Lincoln_Taber_5E4EB9D9EE7E7.pdf (accessed 12.12.13).
Takada, M., Chandra, R.P., Saddler, J.N., 2019. The influence of lignin migration and
relocation during steam pretreatment on the enzymatic hydrolysis of softwood and corn
stover biomass substrates. Biotechnol. Bioeng.1-34. doi:10.1002/bit.27137
Tan, K.T., Lee, K.T., Mohamed, A.R., 2008. Role of energy policy in renewable energy
accomplishment: The case of second-generation bioethanol. Energy Policy 36, 3360–
3365. doi:10.1016/j.enpol.2008.05.016
Thammasittirong, S.N.R., Chatwachirawong, P., Chamduang, T., Thammasittirong, A., 2017.
Evaluation of ethanol production from sugar and lignocellulosic part of energy cane. Ind.
Crops Prod. 108, 598–603. doi:10.1016/j.indcrop.2017.07.023

47
Journal Pre-proof
Tokgoz, S., 2019. Chapter 5 - The food-fuel-fiber debate, in: Debnath, D., Babu, S.C. (Eds.),
Biofuels, Bioenergy and Food Security. Academic Press Massachusetts, US 79–99.
doi.org/10.1016/B978-0-12-803954-0.00005-X
Tomei, J., Helliwell, R., 2016. Food versus fuel? Going beyond biofuels. Land use policy 56,
320–326. doi.org/10.1016/j.landusepol.2015.11.015
Uría-martínez, R., Leiby, P.N., Brown, M.L., 2018. Energy security role of biofuels in
evolving liquid fuel markets. Biofuel, Bioprod. Biorefining 802–814.
doi:10.1002/bbb.1891
van Zyl, W.H., Chimphango, A F. A, den Haan, R., Görgens, J.F., Chirwa, P.W.C., 2011.
Next-generation cellulosic ethanol technologies and their contribution to a sustainable
Africa. Interface Focus 1, 196–211. doi:10.1098/rsfs.2010.0017
Wijaya, H., Sasaki, K., Kahar, P., Yopi, Kawaguchi, H., Sazuka, T., Ogino, C., Prasetya, B.,
Kondo, A., 2018. Repeated ethanol fermentation from membrane-concentrated sweet
sorghum juice using the flocculating yeast Saccharomyces cerevisiae F118 strain.
Bioresour. Technol. 265, 542–547. doi:10.1016/j.biortech.2018.07.039
Wirawan, F., Cheng, C.L., Kao, W.C., Lee, D.J., Chang, J.S., 2012. Cellulosic ethanol
production performance with SSF and SHF processes using immobilized Zymomonas
mobilis. Appl. Energy 100, 19–26. doi:10.1016/j.apenergy.2012.04.032
Wu, L., Li, Y., Arakane, M., Ike, M., Wada, M., Terajima, Y., Ishikawa, S., Tokuyasu, K.,
2011. Efficient conversion of sugarcane stalks into ethanol employing low temperature
alkali pretreatment method. Bioresour. Technol. 102, 11183–8.
doi:10.1016/j.biortech.2011.09.081
Xiros, C., Olsson, L., Koppram, R., Toma, E., 2014. Lignocellulosic ethanol production at
high-gravity : challenges and perspectives. Trends Biotechnol. 32, 36–43.
Yamakawa, C.K., Qin, F., Mussatto, S.I., 2018. Advances and opportunities in biomass
conversion technologies and biore fi neries for the development of a bio-based economy.
Biomass and Bioenergy 119, 54–60. doi:10.1016/j.biombioe.2018.09.007
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. Sustain. Energy Rev. 66, 751–774. doi:10.1016/j.rser.2016.08.038
Zhao, X., Moates, G.K., Wilson, D.R., Ghogare, R.J., Coleman, M.J., Waldron, K.W., 2015.
Steam explosion pretreatment and enzymatic saccharification of duckweed (Lemna
minor) biomass. Biomass and Bioenergy 72, 206–215.
doi:10.1016/j.biombioe.2014.11.003
Zhou, H., Zhan, W., Wang, L., Guo, L., Liu, Y., 2018. Making Sustainable Biofuels and
Sunscreen from Corncobs To Introduce Students to Integrated Biore fi nery Concepts
and Techniques. J. Chem. Educ. 95, 1376–1380. doi:10.1021/acs.jchemed.7b00819

48
Journal Pre-proof

Highlights
 First generation bioethanol production is faced with competing interest of using food-
based feedstocks
 Second generation bioethanol production present the opportunities of using non-food-
based food stocks
 Integration opportunities for First and Second generation bioethanol production are
presented
 The benefits of the different integration scenarios for bioethanol production were
considered
 Future research outlook on the integrated bioethanol production from sugar-based
feedstocks are presented.

You might also like