Renewable Energy Technologies For Energy Efficient Sustainable Development
Renewable Energy Technologies For Energy Efficient Sustainable Development
Arindam Sinharoy
Piet N. L. Lens Editors
Renewable Energy
Technologies
for Energy Efficient
Sustainable
Development
Applied Environmental Science and Engineering
for a Sustainable Future
Series Editors
Jega V. Jegatheesan, School of Engineering, RMIT University, Melbourne, VIC,
Australia
Li Shu, LJS Environment, Melbourne, Australia
Piet N. L. Lens, UNESCO-IHE Institute for Water Education, Delft, The Netherlands
Chart Chiemchaisri, Kasetsart University, Bangkok, Thailand
Applied Environmental Science and Engineering for a Sustainable Future (AESE)
series covers a variety of environmental issues and how they could be solved through
innovations in science and engineering. Our societies thrive on the advancements in
science and technology which pave the way for better standard of living. The adverse
effect of such improvements is the deterioration of the environment. Thus, better
catchment management in order to sustainably manage all types of resources
(including water, minerals and others) is of paramount importance. Water and
wastewater treatment and reuse, solid and hazardous waste management, industrial
waste minimisation, soil restoration and agriculture as well as a myriad of other
topics need better understanding and application. This book series aims at fulfilling
such a task in coming years.
Renewable Energy
Technologies for Energy
Efficient Sustainable
Development
Editors
Arindam Sinharoy Piet N. L. Lens
Department of Microbiology Department of Microbiology
National University of Ireland National University of Ireland
Galway, Ireland Galway, Ireland
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by
similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Contents
Part II Pretreatment
3 Adsorbents for the Detoxification of Lignocellulosic Wastes
Hydrolysates to Improve Fermentative Processes to Bioenergy
and Biochemicals Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Itzel Covarrubias-García and Sonia Arriaga
4 Pretreatment of Lignocellulosic Materials to Enhance their
Methane Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
A. Oliva, S. Papirio, G. Esposito, and P. N. L. Lens
5 Biogas Production from Dairy Cattle Residues: Definition
of the Pretreatment Approach Through a Bibliometric Analysis
of Publications and Patents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Ricardo Müller, Marcio Antonio Vilas Boas, Mônica Sarolli S. M.
Costa, Felipe Souza Marques, Douglas Alves Santos,
Günther Bochmann, Marcelo Bevilacqua Remor,
and Daiana Gotardo Martinez
Abstract The need for a change in the global energy matrix from a fossil fuel based
to a renewable energy one is critical for sustainable development. In this context,
biofuels from anaerobic digestion of agro-industrial waste and wastewater, i.e.,
biohydrogen and biomethane, represent an attractive option. The dark fermentation
process for biohydrogen production involves several possible microbial pathways
that are dynamics and need to be understood to overcome limitations and process
optimization. Methanogenesis for biomethane production occurs close to thermody-
namic limitations. Thus, the system needs to be balanced to achieve stability and
satisfactory biofuel production. Several operational parameters interfere with the
process, and the knowledge about them allows to maximize the methane yield.
Besides problems with gas supersaturation, the role of micronutrients and adequate
removal of toxic compounds released during the biomass decay represent themes
that still must be solved to achieve the full potential of biofuels generation. This
chapter provides fundamental knowledge to overcome the technological limitation
and advance towards a wide production of biofuels from organic waste using
anaerobic digestion.
1.1 Introduction
The growing concern on climate change has driven attention to alternative energy
supplies to convert the current energy matrix based on fossil fuels to renewable ones.
The world’s energy consumption from fossil fuels, i.e., oil, gas, and coal, represented
Fig. 1.1 Global consumption by source in the last 50 years (Data source: Ritchie and Roser 2020)
84.3% in 2019 (Ritchie and Roser 2020), compared with 4.3% from nuclear and
11.4% from renewable sources. A decrease of only 9.4% in the last 50 years
(Fig. 1.1) demonstrating the urge to improve the sustainable energy grid.
Among the renewable sources, i.e., wind, solar, and hydropower, biofuels have
received increasing attention in the last decades. Biofuels can be produced by
microbial fermentation of biomass and presented as solid, liquid, and gaseous
fuels, such as ethanol produced by yeast from sugarcane and corn, biodiesel from
microalgae, and biomethane upgraded biogas from anaerobic fermentation of waste
and wastewater by syntrophic interaction between bacteria and archaea.
CH4 produced from organic matter degradation through anaerobic digestion
(AD) can be upgraded to biomethane (BioCH4) by removing CO2, achieving
>96% purity of CH4 and an energy density of 50–55 kJ g1 (Beil and Beyrich
2013). The AD process occurs in a closed reactor in the absence of oxygen. It is
divided into (1) hydrolysis of macromolecules, (2) fermentation of monomers into
organic acids, hydrogen (H2), CO2 and alcohols, (3) conversion of previous inter-
mediate products into acetate, and (4) final conversion of acetate and hydrogen to
biogas (mixture of CH4 and CO2).
Hydrogen thus produced, also called BioH2, can also be considered a biofuel, as it
is a clean energy carrier with an energy density of 122 kJ g1, 2.75-fold higher than
that of fossil fuels and generating only water when converted to electricity in fuel
cells. The BioH2 production through anaerobic fermentation of organic matter is a
process called dark fermentation. BioH2 is generated to balance the cell redox
potential during fermentation by using the electrons eliminated in the process for
reducing protons and then excreted from the cells as dissolved gas (Cabrol et al.
2017; Schwartz and Friedrich 2006).
1 Fundamentals of Biofuel Production Using Anaerobic Digestion: Metabolic. . . 5
As the dark fermentation process corresponds to the acidogenic step of AD, it can
be followed by acetogenesis and methanogenesis. The latter results in the formation
of biogas, from which the CH4 can be enriched to BioCH4. Arreola-Vargas et al.
(2016) demonstrated that the overall energy recovery from bagasse is higher in a
system performing BioH2 followed by BioCH4 production (two-phase) than a
system producing only BioCH4 (single-phase). Therefore, coupling both processes
can improve the biofuel production, representing advances towards renewable and
sustainable energy production. Therefore, its fundamentals and process interferences
will be addressed in this chapter.
Fig. 1.2 Main metabolic pathways and products associated with dark fermentation from glucose,
considering the Embden–Meyerhof–Parnas (EMP) glycolysis. Pyruvate can be converted into
Acetyl-CoA either through the pyruvate-formate-lyase (PFL) or pyruvate ferredoxin
oxydoreductase (PFOR) pathway depending if the microorganisms are facultative or strict anaer-
obes. In the PFL pathway, the BioH2 and CO2 production from formate might follow a reaction
catalyzed by formate hydrogen lyase (FHL) along with either a [NiFe] Ech (Escherichia coli
hydrogenase 3) hydrogenase or a formate-dependent (COO) [FeFe] hydrogenase. In the PFOR
pathway, a ferredoxin-dependent (Fd-[FeFe]) hydrogenase is involved in the production of BioH2
via NADH: ferredoxin oxidoreductase (NFOR) or NADP+ reducing factor (NRF). The acetyl CoA
branch generates organic acids and alcohol. Pyruvate can directly generate L-lactate, which might
lead to propionate for reducing NADH. Propionate can also be formed through the dicarboxylic acid
(DCA) cycle. Numbers between parentheses represent stoichiometry for 1 mol of glucose. Adapted
from Cabrol et al. (2017) and Hoelzle et al. (2014)
∘
C6 H 12 O6 þ 2H 2 O ! 2 CH 3 COOH þ 4H 2 þ 2CO2 ΔG
206:0 kJ mol1 ð1:1Þ
Acetate and butyrate are the primary pathways in the acetyl Co-A branch, as these
products result in the highest ATP production (Hoelzle et al. 2014). Butyrate is
1 Fundamentals of Biofuel Production Using Anaerobic Digestion: Metabolic. . . 7
Propionate can also be generated by the dicarboxylic acid (DCA) cycle carried
out by Propionibacterium species. The cycle begins when the oxaloacetate is
reduced to L-malate, followed by fumarate and succinate production, which gener-
ates propionate when succinate receives the CoA from the propanoyl-CoA (Hoelzle
et al. 2014). The succinyl-CoA leads to the regeneration of oxaloacetate by trans-
ferring CoA to pyruvate to form propanoyl-CoA (Hoelzle et al. 2014).
reductive direction for CO2 fixation (Ragsdale and Pierce 2008). This pathway is
known as the homoacetogenic pathway and responsible for depletion in BioH2
production, and it is also spread among Clostridium genus (Castelló et al. 2020).
The CH4 production can occur in a single-phase reactor using raw waste or
wastewater or using a two-phase system that receives the effluent of the dark
fermentation reactor. In two-phase systems, the methanogenic reactor is fed with a
substrate formed by a mixture of organic acids and alcohols. Therefore, it is
necessary to convert these compounds into the substrates required by methanogenic
archaea, mainly acetate, CO2 and H2.
Generally, the acetogenic reactions are thermodynamically not favourable, as no
energy is released (Table 1.1). Nevertheless, the consumption of the dissolved
hydrogen and acetate by methanogenic archaea allows a syntrophic relationship
that alters the energetic balance, allowing sufficient energy production to allow the
growth of the acetogenic microorganisms (Madigan et al. 2010).
The global reaction (Eq. 1.13) demonstrates that the propionate yields 56.6 kJ per
mol converted, divided among the groups involved (acetoclastic bacteria,
hydrogenotrophic and acetoclastic methanogens), hence explaining their low growth
rate and biomass yield.
Harper and Pohland (1986) used the results from thermodynamic calculations of
main anaerobic reactions versus hydrogen partial pressure (Fig. 1.3) to establish a
methanogenic niche, in which the hydrogen partial pressure ranges from 106 to
104 atm. This range is thermodynamically favourable for the syntrophism between
organic acid producers and consumers, consequently, for CH4 production.
Operational failures, such as shock loads, biomass wash-out, changes in temper-
ature and pH, might lead to BioH2 and acetate accumulation and inhibiting the
degradation of reduced organic acids, as acetogenic reactions occur close to the
thermodynamic equilibrium. Proprionate concentrations exceeding 900 mg L1
decrease the acetogenic bacterial activity, hampering other organic acids degradation
and leading to more accumulation in the system (Wang et al. 2009). The organic
acids accumulation causes the drop of the system pH, as the alkalinity of the medium
is consumed, hindering methanogenic archaea growth and activity, which causes
acetate accumulation until the CH4 is completely ceased (Wang et al. 2009).
Therefore, methanogenic systems are more sensitive to variations and susceptible
to failures.
Table 1.1 Main reactions involved in acetate and methane production from organic waste and wastewater (Harper and Pohland 1986; Thauer et al. 1977)
Stage Reaction ΔG (kJ) Eq.
Acetogenesis þ +76.1 (1.3)
CH3 CH2 COO þ 3H2 O ! CH3 COO þ HCO
þ H þ 3H2
3
propionate acetate
CH3 CH2 COO þ 2HCO3 ! CH3 COO þ Hþ þ 3HCOO +72.2 (1.4)
propionate acetate
CH3 CH2 CH2 COO þ 2H2 O ! 2CH3 COO þ Hþ þ 2H2 +48.1 (1.5)
butyrate acetate
CH3 CH2 OH þ H2 O ! CH3 COO þ Hþ þ 2H2 +9.6 (1.6)
ethanol acetate
þ 4.2 (1.7)
CH3 CHOHCOO þ 2H2 O ! CH3 COO þ HCO 3 þ H þ 2H2
lactate acetate
þ (1.8)
2HCO3 þ 4H 2 þ H ! CH 3 COO þ 4H 2 O 104.6
bicarbonate acetate
Methanogenesis CH3 COO þ H2 O ! CH4 þ HCO 3 31 (1.9)
acetate methane
4H2 þ CO2 ! CH4 þ 2 H2 O 131 (1.10)
hydrogen methane
HCOO þ ¼ H2 O þ ¼ Hþ ! ¼ CH4 þ ¾ HCO 3 32.6 (1.11)
formate methane
1 Fundamentals of Biofuel Production Using Anaerobic Digestion: Metabolic. . .
þ (1.12)
HCO 3 þ 4H2 þ H ! CH4 þ 3H2 O 135.6
bicarbonate methane
9
10 A. Ferreira Maluf Braga and M. Zaiat
Fig. 1.3 Diagram indicating the favourable niche for BioCH4 production from anaerobic degra-
dation of organic waste and wastewater, based on the hydrogen-dependent thermodynamics of the
main reactions: (1) Propionic acid oxidation to acetic acid. (2) Butyric acid oxidation to acetic acid.
(3) Ethanol to acetic acid. (4) Lactic acid to acetic acid. (5) Acetogenic respiration of bicarbonate.
(6) Methanogenic respiration of bicarbonate. (7) Methanogenic cleavage of acetic acid. Acetic acid,
25 mM; propionic, butyric, lactic acids, and ethanol, 10 mM; bicarbonate, 20 mM; methane 0.7 atm.
(Source: Harper and Pohland 1986)
The microorganisms established in the system are the most crucial factor affecting
the metabolic pathway, and consequently, the biofuel production and yield of the
process. Therefore, the inoculum source plays the primary role.
A pure culture can be selected and used to ensure higher control of a specific step
of the biofuel production through the desired pathway. Microorganisms have an
optimal condition for growth and metabolic activity. Hence, a pure culture allows
product yield maximization during the industrial process. The enhancement of the
BioH2 production rate from monomers, i.e., glucose, by Thermotoga neapolitana
cf. Capnolactica was demonstrate via heterofermentation pathway (acetate and
lactate) varying the inoculum concentration by Dreschke et al. (2018). A
co-culture of isolated strains can collaborate to overcome technical aspects of the
conversion of substrates in biofuel. The combination of pure cultures strictly anaer-
obic and facultative can eliminate O2 in the medium; strains able to hydrolyse
cellulose can provided carbohydrate monomers for high BioH2 producers in dark
fermentation; strains able to produce BioH2 at different pH can maintain the pro-
duction without require buffering (Elsharnouby et al. 2013). A co-culture of E. coli
CECT432 strain and Enterobacter spH1 showed a three-fold higher H2 productivity
1 Fundamentals of Biofuel Production Using Anaerobic Digestion: Metabolic. . . 11
from pure glycerol compared with pure of E. coli CECT432 strain (Maru et al. 2016).
However, isolated cultures require sterilization before using equipment and feeding
the system, turning this option infeasible for real substrates, such as wastewater
containing indigenous microorganisms.
Diverse mixed consortia obtained from indigenous/autochthonous substrate and
inoculum provide the system with notable robustness regarding operational changes,
such as variation in substrate composition and characteristics due the higher micro-
bial diversity without the cost for sterilizing the system and substrate (Cabrol et al.
2017). However, mixed non-isolated consortia for BioH2 production must avoid the
presence of consuming H2 populations, such as methanogenic archaea and sulfate-
reducing bacteria. Therefore, inoculum pre-treatment is an option to select a group of
microorganisms. Thermal (temperatures from 90 to 100 C), acidic (pH lower than
3), and alkaline (pH higher than 11) pre-treatment or a combination of them are
widely adopted to select acidogenic bacteria for dark fermentation and to avoid the
presence of methanogenic microorganisms in the dark fermentation reactor. Some
acidogenic bacteria can form spores under extreme conditions, while archaea typi-
cally present in anaerobic sludge are sensitive to the pre-treatment conditions.
Mockaitis et al. (2020) applied acidic, thermal, and acidic-thermal and thermal-
acidic pre-treatments to a mixed-culture for BioH2 production from xylose. The
authors obtained better results with acidic pre-treatment, achieving a BioH2 yield
equal to 1.57 mol H2 mol xylose1. Rafieenia et al. (2018) evaluated inoculum
pre-treatment for BioH2, followed by BioCH4. Heat shock, aeration, alkaline and an
innovative pre-treatment with saponified frying oil (WFO) were evaluated using
synthetic food waste as the substrate. The authors found better results using the WFO
for BioH2 and BioCH4, respectively, 76.1 mL gVS1 and 598.2 mL gVS1.
Penteado et al. (2013) compared BioH2 production in packet-bed reactor using
pre-treated inoculum from an UASB reactor treating poultry slaughterhouse waste-
water, an UASB fed with swine wastewater and natural fermentation of the substrate.
These authors found higher BioH2 yield using the inoculum from natural fermenta-
tion (2.1 1.8 molH2 molsucrose 1) and higher stability using inoculum pre-treated
with an acid shock (2.0 1.1 molH2 molsucrose 1).
Natural fermentation was also applied using sugarcane vinasse from ethanol and
sugar mill using packed bed (Ferraz et al. 2014) and structured bed reactor (Fuess
et al. 2019) continuous reactors. This inoculation strategy exploits the development
of an adapted consortium in the substrate exposed to the ambient conditions, which
will be further recirculated in the reactor for microorganism attachment (Leite et al.
2008). Autochthonous mixed consortia developed naturally in the substrate such as
mushroom farm waste (Lin et al. 2017) and banana waste (Mazareli et al. 2020) were
also used for BioH2 production in batch experiments and represent a suitable
inoculation approach for BioH2 production from real waste and wastewater.
12 A. Ferreira Maluf Braga and M. Zaiat
1.3.2 pH
1.3.3 Temperature
Fig. 1.4 Solubility of H2, CH4, and CO2 in water. The measurements are normalized for standard
conditions (0 C and 1 atm). (Data source: Dean 1999)
at a certain temperature range. At the lower range value, the semifluid cytoplasmic
membrane begins a solidification process, hindering the transport in the microbial
cell (Madigan et al. 2010). The upper range value indicates the temperature triggers
the collapse of the cytoplasmic membrane, with enzyme denaturation, and thermal
lysis (Madigan et al. 2010). Psychrophilic microorganisms are adapted to tempera-
tures ranging from 0 to 10 C; mesophiles possess optimal temperature at 40 C, but
their metabolism is functioning from 10 to 45 C; thermophiles have an optimal
temperature around 60 C, with activity in the range from 45 to 65 C; and
hyperthermophiles are able to live in environments with temperatures from 90 to
115 C, with an optimum at 105 C (Madigan et al. 2010).
The temperature also affects the solubility of the main gases participating in
AD. The solubility values decrease with the increase in temperature (Fig. 1.4),
indicating that higher operational temperatures might help overcome mass transport
and thermodynamic limitations.
Regarding dark fermentation, the lower solubility of H2 increases the BioH2 yield
by enhancing the mass transfer from the liquid to the gaseous phase and limiting the
homoacetogenesis due to the lower CO2 solubility. In methanogenesis, the lower
solubility of CH4 allows the higher mass transfer of CH4 to the gaseous phase.
However, the lower CO2 solubility might impact the activity of hydrogenotrophic
methanogens. Labatut et al. (2014) found that mesophilic continuously-stirred
anaerobic digesters (CSADs) producing CH4 from cow manure co-digested with
dog fod waste was more robust and stable than a thermophilic CSAD, despite a
slightly lower CH4 yield.
Supersaturation of CH4 in the liquid phase has already been reported by Yeo et al.
(2015). Despite the low solubility of H2 (<CH4 << CO2), BioH2 supersaturation
was observed in several studies under mesophilic and thermophilic conditions
(Beckers et al. 2015; Dreschke et al. 2019a, b; Obazu et al. 2012; Zhang et al.
14 A. Ferreira Maluf Braga and M. Zaiat
2012). Values up to seven-fold the values calculated for equilibrium conditions were
found by Beckers et al. (2015) during BioH2 production using the pure Clostridium
butyricum strain, demonstrating that other strategies besides increasing temperature
must be applied to promote a higher BioH2 yield.
lignocellulosic material type and the pre-treatment technique applied (Monlau et al.
2014).
The BioH2 production can be inhibited by furanic and phenolic compounds with
a concentration range of 250–1000 mg L1 (Elbeshbishy et al. 2017). The inhibition
by phenolic compounds is related to the damage in the microbial cell membrane,
while the furans interfere with enzyme activity (Elbeshbishy et al. 2017). Fang et al.
(2006) demonstrated 70% phenol degradation at a concentration of 630 mg L1 in a
thermophilic UASB at HRT of 28 h. A specific methanogenic activity (SMA) test
demonstrated that the sludge was able to degrade the phenols at a range concentra-
tion of 600–1000 mg L1 and produce CH4; however, no phenol degradation was
achieved at 2000 mg L1 (Fang et al. 2006).
Ideally, the liquid flow pattern in continuous reactors can be classified in plug-flow
and mixed flow, determined by models such as tank-in-series and dispersion
(Levenspiel 1999). It directly impacts the kinetics of the process. Some reactor
configurations applied to biofuels production are presented in Table 1.2. The
mixed flow present in the continuous stirred tank reactor (CSTR) allows a homog-
enous composition throughout the reactor and favours the mass transfer of the
products excreted by the microorganisms (Bailey and Ollis 1986). The plug-flow
pattern allows a longitudinal microorganism separation, and consequently, the
composition of the medium changes throughout the reactor, and it is suitable for
substrates with potentially toxic compounds (Parkin and Speece 1983). Plug-flow
reactors can be used in large-scale for CH4 production from cattle manure in long
term operation (Dong et al. 2019). However, most of the reactors possess a non-ideal
flow pattern, standing in between mixed and plug-flow.
Important reactors characteristics are the presence or absence of a support mate-
rial for microbial attachment, recirculation of liquid or gas, and up or downflow
velocity. Reactor modifications that promote sludge settling and aggregation of
biomass, e.g. three-phase separator in UASB reactors, disconnect the hydraulic
retention time (HRT) from the sludge retention time (SRT), allowing a high-rate
reactor (Lettinga et al. 1980). This also occurs when adding support material such as
in packed and structured bed reactors. Liquid recirculation can be used to increase
the upflow velocities as in expanded granular sludge blanket (EGSB) reactors,
increasing the contact between substrate and microorganisms (Verstraete et al.
1996). It can also improve the alkalinity in methanogenic reactors and dilute toxic
compounds present in the system. Gas recirculation can be applied to overcome H2
supersaturation (Beckers et al. 2015; Dreschke et al. 2019a).
Table 1.2 Reactors used for biofuels production through anaerobic digestion
16
ORL
(gCOD Temperature
Stage Reactor Substrate Inoculum L1 d1) ( C) pH initial Biofuel yield Biofuel production Reference
Dark Continuous Glucose Thermotoga – 80 7 3.5 0.2 molH2 850 71 mLH2 Dreschke
fermentation stirred tank neapolitana mol1 L1 h1 et al.
reactor cf. capnolactica (2019b)
(CSTR)
Dark Internal cir- Glucose Mixed culture – 35 1 6.5 1.4 0.21 molH2 10.78 0.84 L1 Su et al.
fermentation culation (thermal mol1 d1 (2020)
(IC) reactor pre-treatment)
Dark CSTR Tequila Vinasse Mixed culture 32.8 37 7.5 1.68 molH2 mol1 0.42 0.02 NL H2 García-
fermentation (thermal L1 d1 Becerra
pre-treatment) et al.
(2019)
Dark Up-flow Glucose Mixed culture 64 48 2 6.0 0.5 0.89 molH2 mol1 4.73 LH2 L1 d1 Karapinar
fermentation packed bed (thermal et al.
reactor pre-treatment) (2020)
(UPBR)
Dark Anaerobic Sugarcane vinasse Natural 56 70 6.5 2.6 0.5 molH2 630 47 mLH2 Niz et al.
fermentation structured fermentation mol1 L1 d1 (2019)
bed reactor
(AnSTBR)
Methanogenesis Plug-flow Cattle manure Mixed culture 3.6 37–40 7.85 0.05 – 0.43–0.75m3 m3 Dong
reactor d1 et al.
(2019)
Methanogenesis Fixed-bed Municipal organic Mixed culture 0.7 37–40 7–8 – 335 N-LCH4 Knoop
digester waste percolate kgVS-1 a et al.
(two-phase (2018)
system)
Methanogenesis Horizontal- Cassava Mixed culture 2.3–8.5 28.5 1.5 7-8b 0.3–0.15 LCH4 0.5–4.5 LCH4 d1 Palma
flow reac- processing waste- gCODremoved1 et al.
tors with water CPWW (2016)
A. Ferreira Maluf Braga and M. Zaiat
dolomite
rocks
Methanogenesis CSTR Sugarcane trash Mixed culture 2.5 (gVS 37 1 7.3–7.8b 121 (ST), 303 (ST), Paulose
untreated (ST) and L1 d1)c 226 (TT), 565 (TT), and
treated 148 (SB), and 370 (SB) and Kaparaju
(TT) sugarcane 236 (TB) (mLCH4 590 (TB) (mLCH4 (2021)
bagasse untreated gVSfed1) L1 d1)
(SB) and treated
(TB)
Methanogenesis CSTR and Swine manure Mixed culture 0.165 lab 23–24 BLC 7.4 BLC 0.9 lab and 0.56 0.2 lab and 0.18 Tápparo
Covered solid (CSTR) and and 0.356 34–37 CSTR and 7.8 full CLB, 0.46 lab full CLB, 0.61 lab et al.
lagoon liquid (CLB) full CLB; CSTR and 0.38 full CSTR and 0.65 full CSTR (2021)
biodigester phase 1.340 lab (NLbiogas gVSadd1 (NLbiogas L1 d1)
(CLB) lab and 1.618
and full full CSTR
scale
a
Value from biochemical potential assay
b
Effluent pH
c
Best OLR
1 Fundamentals of Biofuel Production Using Anaerobic Digestion: Metabolic. . .
17
18 A. Ferreira Maluf Braga and M. Zaiat
References
Angenent LT, Karim K, Al-Dahhan MH, Wrenn, B. a, & Domíguez-Espinosa, R. (2004) Produc-
tion of bioenergy and biochemicals from industrial and agricultural wastewater. Trends
Biotechnol 22(9):477–485. https://doi.org/10.1016/j.tibtech.2004.07.001
Arreola-Vargas J, Flores-Larios A, González-Álvarez V, Corona-González RI, Méndez-Acosta HO
(2016) Single and two-stage anaerobic digestion for hydrogen and methane production from
acid and enzymatic hydrolysates of Agave tequilana bagasse. Int J Hydrog Energy
41(2):897–904. https://doi.org/10.1016/j.ijhydene.2015.11.016
Bailey JE, Ollis DF (1986) Biochemical engineering fundamentals, 2nd edn. McGraw-Hill,
New York
Beckers L, Masset J, Hamilton C, Delvigne F, Toye D, Crine M, Thonart P, Hiligsmann S (2015)
Investigation of the links between mass transfer conditions, dissolved hydrogen concentration
and biohydrogen production by the pure strain Clostridium butyricum CWBI1009. Biochem
Eng J 98:18–28. https://doi.org/10.1016/j.bej.2015.01.008
Beil M, Beyrich W (2013) Biogas upgrading to biomethane. In: Wellinger A, Murphy JD, Baxter D
(eds) The biogas handbook: science, production and applications. Woodhead, Cambridge, pp
342–377
Cabrol L, Marone A, Tapia-Venegas E, Steyer JP, Ruiz-Filippi G, Trably E (2017) Microbial
ecology of fermentative hydrogen producing bioprocesses: useful insights for driving the
ecosystem function. FEMS Microbiol Rev 41(2):158–181. https://doi.org/10.1093/femsre/
fuw043
Cai G, Jin B, Monis P, Saint C (2011) Metabolic flux network and analysis of fermentative
hydrogen production. Biotechnol Adv 29(4):375–387. https://doi.org/10.1016/j.biotechadv.
2011.02.001
Can M, Armstrong FA, Ragsdale SW (2014) Structure, function, and mechanism of the nickel
metalloenzymes, CO dehydrogenase, and acetyl-CoA synthase. Chem Rev 114(8):4149–4174.
https://doi.org/10.1021/cr400461p
Castelló E, Nunes Ferraz-Junior AD, Andreani C, del Anzola-Rojas MP, Borzacconi L, Buitrón G,
Carrillo-Reyes J, Gomes SD, Maintinguer SI, Moreno-Andrade I, Palomo-Briones R, Razo-
Flores E, Schiappacasse-Dasati M, Tapia-Venegas E, Valdez-Vázquez I, Vesga-Baron A,
Zaiat M, Etchebehere C (2020) Stability problems in the hydrogen production by dark fermen-
tation: possible causes and solutions. Renew Sust Energ Rev 119(November 2019):109602.
https://doi.org/10.1016/j.rser.2019.109602
Chabrière E, Charon MH, Volbeda A, Pieulle L, Hatchikian EC, Fontecilla-Camps JC (1999)
Crystal structures of the key anaerobic enzyme pyruvate ferredoxin oxidoreductase free and in
complex with pyruvate. Nat Struct Biol 6(2):182–190. https://doi.org/10.1038/5870
Dean JA (1999) Lange’s handbook of chemistry, 15th edn. McGraw-Hill. https://www.amazon.
com/Langes-Handbook-Chemistry-John-Dean/dp/0070163847
Dong L, Cao G, Guo X, Liu T, Wu J, Ren N (2019) Efficient biogas production from cattle manure
in a plug flow reactor: A large scale long term study. Bioresour Technol 278:450–455. https://
doi.org/10.1016/j.biortech.2019.01.100
Dreschke G, d’Ippolito G, Panico A, Lens PNL, Esposito G, Fontana A (2018) Enhancement of
hydrogen production rate by high biomass concentrations of Thermotoga neapolitana. Int J
Hydrog Energy 43(29):13072–13080. https://doi.org/10.1016/j.ijhydene.2018.05.072
Dreschke G, Papirio S, d’Ippolito G, Panico A, Lens PNL, Esposito G, Fontana A (2019a) H2-rich
biogas recirculation prevents hydrogen supersaturation and enhances hydrogen production by
Thermotoga neapolitana cf. capnolactica. Int J Hydrog Energy 44:19698–19708. https://doi.
org/10.1016/j.ijhydene.2019.06.022
Dreschke G, Papirio S, Lens PNL, Esposito G (2019b) Influence of liquid-phase hydrogen on dark
fermentation by Thermotoga neapolitana. Renew Energy 140:354–360. https://doi.org/10.
1016/j.renene.2019.02.126
1 Fundamentals of Biofuel Production Using Anaerobic Digestion: Metabolic. . . 19
Elbeshbishy E, Dhar BR, Nakhla G, Lee HS (2017) A critical review on inhibition of dark
biohydrogen fermentation. Renew Sust Energ Rev 79(October 2015):656–668. https://doi.org/
10.1016/j.rser.2017.05.075
Elsharnouby O, Hafez H, Nakhla G, El Naggar MH (2013) A critical literature review on
biohydrogen production by pure cultures. Int J Hydrog Energy 38(12):4945–4966. https://doi.
org/10.1016/j.ijhydene.2013.02.032Fang
Fang HHP, Liang DW, Zhang T, Liu Y (2006) Anaerobic treatment of phenol in wastewater under
thermophilic condition. Water Res 40(3):427–434. https://doi.org/10.1016/j.watres.2005.
11.025
Ferraz ADN, Wenzel J, Etchebehere C, Zaiat M (2014) Effect of organic loading rate on hydrogen
production from sugarcane vinasse in thermophilic acidogenic packed bed reactors. Int J Hydrog
Energy 39(30):16852–16862. https://doi.org/10.1016/j.ijhydene.2014.08.017
Fuess LT, Zaiat M, do Nascimento, C. A. O. (2019) Novel insights on the versatility of biohydrogen
production from sugarcane vinasse via thermophilic dark fermentation: impacts of pH-driven
operating strategies on acidogenesis metabolite profiles. Bioresour Technol 286:121379. https://
doi.org/10.1016/j.biortech.2019.121379
Garcia C, Molina F, Roca E, Lema JM (2007) Fuzzy-based control of an anaerobic reactor treating
wastewaters containing ethanol and carbohydrates. Ind Eng Chem Res 46(21):6707–6715.
https://doi.org/10.1021/ie0617001
García-Becerra M, Macías-Muro M, Arellano-García L, Aguilar-Juárez O (2019) Bio-hydrogen
production from tequila vinasses: effect of detoxification with activated charcoal on dark
fermentation performance. Int J Hydrog Energy 44(60):31860–31872. https://doi.org/10.1016/
j.ijhydene.2019.10.059
Harper SR, Pohland FG (1986) Recent developments in hydrogen management during anaerobic
biological wastewater treatment. Biotechnol Bioeng 28(4):585–602. https://doi.org/10.1002/bit.
260280416
Hoelzle RD, Virdis B, Batstone DJ (2014) Regulation mechanisms in mixed and pure culture
microbial fermentation. Biotechnol Bioeng 111(11):2139–2154. https://doi.org/10.1002/bit.
25321
Horiuchi JI, Shimizu T, Tada K, Kanno T, Kobayashi M (2002) Selective production of organic
acids in anaerobic acid reactor by pH control. Bioresour Technol 82(3):209–213
Hwang J-H, Choi J-A, Abou-Shanab RAI, Bhatnagar A, Min B, Song H, Kumar E, Choi J, Lee ES,
Kim YJ (2009) Effect of pH and sulfate concentration on hydrogen production using anaerobic
mixed microflora. Int J Hydrog Energy 34(24):9702–9710. https://doi.org/10.1016/j.ijhydene.
2009.10.022
Karapinar I, Gokfiliz Yildiz P, Pamuk RT, Karaosmanoglu Gorgec F (2020) The effect of hydraulic
retention time on thermophilic dark fermentative biohydrogen production in the continuously
operated packed bed bioreactor. Int J Hydrog Energy 45(5):3524–3531. https://doi.org/10.1016/
j.ijhydene.2018.12.195
Kiyuna LSM, Fuess LT, Zaiat M (2017) Unraveling the influence of the COD/sulfate ratio on
organic matter removal and methane production from the biodigestion of sugarcane vinasse.
Bioresour Technol 232:103–112. https://doi.org/10.1016/j.biortech.2017.02.028
Knoop C, Tietze M, Dornack C, Raab T (2018) Fate of nutrients and heavy metals during two-stage
digestion and aerobic post-treatment of municipal organic waste. Bioresour Technol 251:238–
248. https://doi.org/10.1016/j.biortech.2017.12.019
Labatut RA, Angenent LT, Scott NR (2014) Conventional mesophilic vs. thermophilic anaerobic
digestion: Atrade-off between performance and stability? Water Res 53:249–258
Leite JAC, Fernandes BS, Pozzi E, Barboza M, Zaiat M (2008) Application of an anaerobic packed-
bed bioreactor for the production of hydrogen and organic acids. Int J Hydrog Energy
33(2):579–586. https://doi.org/10.1016/j.ijhydene.2007.10.009
Lettinga G, van Velsen AFM, Hobma SW, de Zeeuw W, Klapwijk A (1980) Use of the upflow
sludge blanket (USB) reactor concept for biological wastewater treatment, especially for
anaerobic treatment. Biotechnol Bioeng 22(4):699–734. https://doi.org/10.1002/bit.260220402
20 A. Ferreira Maluf Braga and M. Zaiat
Levenspiel O (1999) Chemical reaction engineering, 3rd edn. Wiley, New York. https://doi.org/10.
1016/0009-2509(64)85017-X
Lin CY, Lay CH, Sung IY, Sen B, Chen CC (2017) Anaerobic hydrogen production from
unhydrolyzed mushroom farm waste by indigenous microbiota. J Biosci Bioeng 124:425–
429. https://doi.org/10.1016/j.jbiosc.2017.05.001
Liu T, Sung S (2002) Ammonia inhibition on thermophilic aceticlastic methanogens. Water Sci
Technol 45(10):113–120. https://doi.org/10.2166/wst.2002.0304
Madigan MT, Martinko JM, Stahl DA, Clark DP (2010) Brock biology of microorganisms, 13th
edn. Benjamin Cummings, San Francisco
Maru BT, López F, Kengen SWM, Constantí M, Medina F (2016) Dark fermentative hydrogen and
ethanol production from biodiesel waste glycerol using a co-culture of Escherichia coli and
Enterobacter sp. Fuel 186:375–384. https://doi.org/10.1016/j.fuel.2016.08.043
Mazareli RC d S, Villa-Montoya AC, Delforno TP, Centurion VB, Maia de Oliveira V, Silva EL,
Amâncio Varesche MB (2020) Metagenomic analysis of autochthonous microbial biomass from
banana waste: screening design of factors that affect hydrogen production. Biomass Bioenergy
138. https://doi.org/10.1016/j.biombioe.2020.105573
McDowall JS, Murphy BJ, Haumann M, Palmer T, Armstrong FA, Sargent F (2014) Bacterial
formate hydrogenlyase complex. Proc Natl Acad Sci U S A 111(38):E3948–E3956. https://doi.
org/10.1073/pnas.1407927111
Mockaitis G, Bruant G, Guiot SR et al (2020) Acidic and thermal pre-treatments for anaerobic
digestion inoculum to improve hydrogen and volatile fatty acid production using xylose as the
substrate. Renew Energy 145:1388–1398. https://doi.org/10.1016/j.renene.2019.06.134
Molaey R, Bayrakdar A, Sürmeli RÖ, Çalli B (2019) Anaerobic digestion of chicken manure:
influence of trace element supplementation. Eng Life Sci 19(2):143–150. https://doi.org/10.
1002/elsc.201700201
Monlau F, Sambusiti C, Barakat A, Quéméneur M, Trably E, Steyer JP, Carrère H (2014) Do
furanic and phenolic compounds of lignocellulosic and algae biomass hydrolyzate inhibit
anaerobic mixed cultures? A comprehensive review. Biotechnol Adv 32(5):934–951. https://
doi.org/10.1016/j.biotechadv.2014.04.007
Mulrooney SB, Hausinger RP (2003) Nickel uptake and utilization by microorganisms. FEMS
Microbiol Rev 27(2–3):239–261. https://doi.org/10.1016/S0168-6445(03)00042-1
Niz MYK, Etchelet I, Fuentes L, Etchebehere C, Zaiat M (2019) Extreme thermophilic condition:
An alternative for long-term biohydrogen production from sugarcane vinasse. Int J Hydrog
Energy 44(41):22876–22887. https://doi.org/10.1016/j.ijhydene.2019.07.015
Obazu FO, Ngoma L, Gray VM (2012) Interrelationships between bioreactor volume, effluent
recycle rate, temperature, pH, %H2, hydrogen productivity and hydrogen yield with undefined
bacterial cultures. Int J Hydrog Energy 37(7):5579–5590. https://doi.org/10.1016/j.ijhydene.
2012.01.001
Onumaegbu C, Mooney J, Alaswad A, Olabi AG (2018) Pre-treatment methods for production of
biofuel from microalgae biomass. Renew Sust Energ Rev 93(May):16–26. https://doi.org/10.
1016/j.rser.2018.04.015
Palma D, Fuess LT, de Lima-Model AN, Zanella da Conceição K, Cereda MP, Ferreira Tavares
MH, Gomes SD (2016) Using dolomitic limestone to replace conventional alkalinization in the
biodigestion of rapid acidification cassava processing wastewater. J Clean Prod 172:2942–2953.
https://doi.org/10.1016/j.jclepro.2017.11.118
Parkin GF, Speece RE (1983) Attached versus suspended growth anaerobic reactors: response to
toxic substances. Water Sci Technol 15(8–9):261–289. https://doi.org/10.2166/wst.1983.0171
Paulose P, Kaparaju P (2021) Anaerobic mono-digestion of sugarcane trash and bagasse with and
without pretreatment. Ind Crop Prod 170:113712. https://doi.org/10.1016/j.indcrop.2021.
113712
Penteado ED, Lazaro CZ, Sakamoto IK, Zaiat M (2013) Influence of seed sludge and pretreatment
method on hydrogen production in packed-bed anaerobic reactors. Int J Hydrog Energy
38(14):6137–6145. https://doi.org/10.1016/j.ijhydene.2013.01.067
1 Fundamentals of Biofuel Production Using Anaerobic Digestion: Metabolic. . . 21
Pinske C, Sargent F (2016) Exploring the directionality of Escherichia coli formate hydrogenlyase:
a membrane-bound enzyme capable of fixing carbon dioxide to organic acid. Microbiol Open
5(5):721–737. https://doi.org/10.1002/mbo3.365
Rafieenia R, Pivato A, Lavagnolo MC (2018) Effect of inoculum pre-treatment on mesophilic
hydrogen and methane production from food waste using two-stage anaerobic digestion. Int J
Hydrog Energy 43(27):12013–12022. https://doi.org/10.1016/j.ijhydene.2018.04.170
Ragsdale SW, Pierce E (2008) Acetogenesis and the wood-Ljungdahl pathway of CO2 fixation.
Biochim Biophys Acta Proteins Proteom 1784(12):1873–1898. https://doi.org/10.1016/j.
bbapap.2008.08.012
Ripley LE, Boyle WC, Converse JC (1986) Improved alkalimetric monitoring for anaerobic
digestion of high-strength wastes. J Water Pollut Control Fed 58(5):406–411. http://www.
jstor.org/stable/25042933
Ritchie H, Roser M (2020) Energy. OurWorldInData.Org. https://ourworldindata.org/energy
Saady NMC (2013) Homoacetogenesis during hydrogen production by mixed cultures dark
fermentation: unresolved challenge. Int J Hydrog Energy 38(30):13172–13191. https://doi.
org/10.1016/j.ijhydene.2013.07.122
Schwartz E, Friedrich B (2006) The H2-metabolizing prokaryotes. In: Rosenberg E, DeLong EF,
Lory S, Stackebrandt E, Thompson FL (eds) The prokaryotes. Springer, New York. https://doi.
org/10.1007/0-387-30742-7_17
Sprott GD, Shaw KM, Jarrell KF (1984) Ammonia/potassium exchange in methanogenic bacteria. J
Biol Chem 259(20):12602–12608. https://doi.org/10.1016/s0021-9258(18)90789-1
Su C, Liu Y, Yang X, Li H (2020) Effect of hydraulic retention time on biohydrogen production
from glucose in an internal circulation reactor. Energy Fuels 34(3):3244–3249. https://doi.org/
10.1021/acs.energyfuels.9b03316
Tápparo DC, Cândido D, Steinmetz RLR, Etzkorn C, do Amaral AC, Antes FG, Kunz A (2021)
Swine manure biogas production improvement using pre-treatment strategies: lab-scale studies
and full-scale application. Bioresour Technol Rep 15:1–8. https://doi.org/10.1016/j.biteb.2021.
100716
Thanh PM, Ketheesan B, Yan Z, Stuckey D (2016) Trace metal speciation and bioavailability in
anaerobic digestion: A review. Biotechnol Adv 34(2):122–136. https://doi.org/10.1016/j.
biotechadv.2015.12.006
Thauer RK, Jungermann K, Decker K (1977) Energy conservation in chemotrophic anaerobic
bacteria. Bacteriol Rev 41(1):100–180. https://doi.org/10.1128/br.41.1.100-180.1977
Verstraete W, de Beer D, Pena M, Lettinga G, Lens P (1996) Anaerobic bioprocessing of organic
wastes. World J Microbiol Biotechnol 12(3):221–238. https://doi.org/10.1007/BF00360919
Vignais PM, Colbeau A (2004) Molecular biology of microbial hydrogenases. Curr Issues Mol Biol
6:159–188
Wang Y, Zhang Y, Wang J, Meng L (2009) Effects of volatile fatty acid concentrations on methane
yield and methanogenic bacteria. Biomass Bioenergy 33(5):848–853. https://doi.org/10.1016/j.
biombioe.2009.01.007
Yeo H, An J, Reid R, Rittmann BE, Lee H-S (2015) Contribution of liquid/gas mass-transfer
limitations to dissolved methane oversaturation in anaerobic treatment of dilute wastewater.
Environ Sci Technol 49(17):10366–10372. https://doi.org/10.1021/acs.est.5b02560
Zhang F, Zhang Y, Chen M, Zeng RJ (2012) Hydrogen supersaturation in thermophilic mixed
culture fermentation. Int J Hydrog Energy 37(23):17809–17816. https://doi.org/10.1016/j.
ijhydene.2012.09.019
Zhang W, Chen B, Li A, Zhang L, Li R, Yang T, Xing W (2019) Mechanism of process imbalance
of long-term anaerobic digestion of food waste and role of trace elements in maintaining
anaerobic process stability. Bioresour Technol 275(December 2018):172–182. https://doi.org/
10.1016/j.biortech.2018.12.052
Chapter 2
Engineering Direct Interspecies Electron
Transfer for Enhanced Methanogenic
Performance
Changsoo Lee
Abstract Producing biogas from organic waste streams through anaerobic diges-
tion (AD) is a well-established bioenergy technology. Efficient electron transfer
between syntrophic bacteria and methanogens is critical for balancing acidogenesis
and methanogenesis, which is necessary for stable digester operation. The recently
discovered direct interspecies electron transfer (DIET) links syntrophic partners via
cell-to-cell electrical connections without using diffusive electron carriers such as
H2. Promoting DIET by adding conductive materials has been suggested as a
possible method to accelerate syntrophic degradation of organic compounds, and
many studies have demonstrated the enhancement of methanogenesis by the addition
of conductive materials. Although further research is needed for practical applica-
tions, accumulated evidence indicates that engineering DIET is a promising strategy
to enhance the performance and stability of AD processes. A few recent studies have
also demonstrated the scale-up potential of DIET-aided AD.
2.1 Introduction
C. Lee (*)
Department of Urban and Environmental Engineering, Ulsan National Institute of Science and
Technology (UNIST), Ulsan, Republic of Korea
e-mail: [email protected]
(Fernandez et al. 2000). The microbial community structure and interspecies inter-
actions will be even more complicated in anaerobic digesters treating complex
organic wastes. Therefore, facilitating and harmonizing the activities of individual
community members is the key to stable and robust AD.
Methanogenesis, the last step of AD, is performed by a unique group of strictly
anaerobic microorganisms belonging to the domain Archaea, known as
methanogens. All methanogens identified so far are classified into one phylum,
Euryarchaeota, while the existence of putative methanogens in uncultured phyla
Bathyarchaeota (Evans et al. 2015) and Verstraetearchaeota (Vanwonterghem et al.
2016) was suggested in recent metagenomic studies. Methanogens use a narrow
range of substrates such as H2/CO2, formate, acetate, methylated compounds and
CO, and acetoclastic and hydrogenotrophic pathways are often the main routes of
methanogenesis in anaerobic digesters (Enzmann et al. 2018). Therefore,
methanogens at the end of the trophic chain of AD need other microorganisms
that hydrolyze and ferment organic macromolecules into substrates for
methanogenesis. Methanogenesis is usually considered the rate-limiting step of the
overall AD process because of the slow growth rate and high sensitivity to environ-
mental conditions of methanogens (Lee et al. 2018). However, hydrolysis can be the
rate-limiting step when digesting organic substances with low bioavailability, such
as waste activated sludge and lignocellulosic biomass. Furthermore, the degradation
of C3 or higher volatile fatty acids (VFAs) to acetate and H2/CO2 (i.e., acetogenesis)
also becomes rate-limiting if the syntrophic association between VFA oxidizers and
methanogens is not well developed (Baek et al. 2018). In this case, an imbalance
between the production and consumption of VFAs (and H2) can occur and result in a
buildup of VFAs (and H2 partial pressure) and even in a digester failure. Therefore,
efficient electron transfer between syntrophic microorganisms involved in anaerobic
VFA degradation is critical in maintaining the balance between acidogenesis and
methanogenesis for stable AD.
The interspecies electron transfer (IET) between VFA oxidizers and methanogens
has been thought to be mediated exclusively by microbially produced H2 or formate
as electron carriers. It was recently revealed that there exists an alternative IET route
where electrons are transferred directly from exoelectrogenic VFA oxidizers to
electrotrophic methanogens through a cell-to-cell electrical connection. This direct
IET (DIET) is energetically and kinetically advantageous over the H2/formate-
mediated indirect IET (IIET), because complex reactions for synthesizing and
consuming H2 or formate are not required (Lovley 2011). This suggests the possi-
bility that thermodynamically more favorable conditions for rapid methanogenic
degradation of organic matter can be achieved by promoting DIET (Barua and Dhar
2017). In support of such a possibility, a mathematical modeling study estimated that
the cell-to-cell electron transfer rate can be more than eight-fold higher for DIET
than for IIET via H2 (Storck et al. 2016).
The potential for DIET in methanogenic systems was experimentally observed
for the first time only a decade ago by Morita et al. (2011) in a study of electron
transfer mechanisms within methanogenic granular sludge treating brewery waste.
In the next year, Kato et al. (2012) reported a significant acceleration of
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 25
IIET between VFA oxidizers and methanogens via H2 or formate has been consid-
ered the major route for syntrophic electron exchange necessary for methanogenic
degradation of C3 or higher VFAs (Fig. 2.1). This syntrophic relationship is very
sensitive to the accumulation of oxidation products of VFAs (i.e., H2 and formate)
26 C. Lee
Table 2.1 Reactions involved in syntropic oxidation of propionate and butyrate by interspecies
electron transfer via hydrogen or formate (adapted from Baek et al. 2018)
Reaction ΔG0 a (kJ/mol)
Methanogenic degradation of propionate via H2 as electron carrier
4CH3CH2COO + 12H2O ! 4CH3COO + 4HCO3 + 4H+ + 12H2 +76.5
3HCO3 + 3H+ + 12H2 ! 3CH4 + 9H2O 101.7
(overall) 4CH3CH2COO + 3H2O ! 4CH3COO + HCO3 + H+ + 3CH4 25.2
Methanogenic degradation of propionate via formate as electron carrier
4CH3CH2COO + 8H2O + 8CO2 ! 4CH3COO + 12HCOO + 4H+ +65.3
12HCOO + 12H+ ! 3CH4 + 9CO2 + 6H2O 144.5
(overall) 4CH3CH2COO + 8H+ + 2H2O ! 4CH3COO + CO2 + 3CH4 79.2
Methanogenic degradation of butyrate via H2 as electron carrier
2CH3CH2CH2COO + 4H2O ! 4CH3COO + 2H+ + 4H2 +48.3
HCO3 + H+ + 4H2 ! CH4 + 3H2O 67.8
(overall) 2CH3CH2CH2COO + H2O + HCO3 ! 4CH3COO + H+ + CH4 19.5
Methanogenic degradation of butyrate via formate as electron carrier
2CH3CH2CH2COO + 4H2O + 4CO2 ! 4CH3COO + 4HCOO + 4H+ +38.5
4HCOO + 4H+ ! CH4 + 3CO2 + 2H2O 96.3
(overall) 2CH3CH2CH2COO + 2H2O + CO2 ! 4CH3COO + CH4 57.8
a
ΔG0 , Standard Gibbs free energy change of reaction at pH 7
Fig. 2.1 Indirect interspecies electron transfer (IIET) between syntrophic VFA-oxidizing bacteria
and hydrogenotrophic methanogens via H2 and formate
Fig. 2.2 Diffusive flux of interspecies electron carriers between a producing and a consuming
microorganism (adapted from de Bok et al. 2004). A, the surface area of the electron carrier
producer; D, the diffusion coefficient of the electron carrier; Cp, the concentration of the electron
carrier at the surface of the producer; Cc, the concentration of the electron carrier at the surface of
the consumer; and d, the distance between the producer and consumer
2.2.2.1 Introduction
Recent discoveries revealed that DIET is an important alternative IET route for the
use of electrons derived from the oxidation of acidogenic metabolites, such as VFAs
and ethanol, for the methanogenic reduction of CO2 in AD environments (Morita
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 29
et al. 2011; Kato et al. 2012; Rotaru et al. 2014a, b). In contrast to IIET, DIET
excludes the involvement of electron carriers (i.e., H2 and formate) because electrons
released by exoelectrogenic bacteria flow directly via intercellular electrical connec-
tions to electrotrophic methanogens. Therefore, electrically coupled syntrophic
partners by DIET can avoid complex reactions for the synthesis, diffusion, and
utilization of interspecies electron carriers, which makes DIET energetically and
kinetically advantageous over IIET (Lovley 2011). An imbalance between the
production and consumption of VFAs often occurs in AD processes due to the
large difference in growth rate between acidogens and methanogens, especially
under high organic loading conditions. In such cases, VFAs (and H2 partial pressure)
build up to levels that cannot be effectively stabilized by IIET-driven
hydrogenotrophic methanogenesis, resulting in severe inhibition of methanogenic
activity. DIET-driven electrotrophic methanogenesis can accelerate the stabilization
of VFA accumulation and help to maintain a stable methanogenic system because of
the higher electron transfer efficiency and smaller intermediate loss in DIET com-
pared to IIET (Kato et al. 2012; Leng et al. 2018). For example, the Gibbs free
energy change at pH 7 (ΔG0 ) of the propionate oxidation through DIET (Eq. 2.1) is
significantly lower than that of the propionate oxidation through IIET via H2
(Eq. 2.2), with the former being negative (i.e., spontaneous reaction) and the latter
being positive (i.e., non-spontaneous reaction) (Jing et al. 2017).
The concentration gradient of electron carriers between syntrophic partners is the
main limiting factor for IIET, while that for DIET is considered to be the activation
loss (or overpotential) during the flow of electrons through cellular components
comprising cell-to-cell electrical connections (Storck et al. 2016). Activation loss is
the energy lost to overcome the energy barrier and initiate electron transfer, for
example, between terminal membrane-bound redox proteins and e-pili in DIET. It
takes place as a result of slow charge transfer in an electrochemical redox reaction,
and a certain portion of the electron energy is consumed to drive the reaction at a
certain rate (Petrovic 2021).
2014b; Rotaru et al. 2014a, b; Shrestha et al. 2014). Given the compact and dense
structure of granular sludge, DIET can explain the efficient degradation of propio-
nate by anaerobic granules even when the Gibbs free energy change calculated from
the dissolved H2 available for the granules was positive (Dubé and Guiot 2015).
DIET mechanisms can be classified into two types: biological DIET (bDIET) and
mineral DIET (mDIET). The former involves direct cell-to-cell contact via biotic
components, whereas abiotic conductive materials support electrical connections
between cells in the latter (Shrestha and Rotaru 2014). Electrons are transferred via
electron transport proteins, such as outer surface c-type cytochromes, or electrically
conductive pili (e-pili), also known as nanowires, in bDIET (Fig. 2.3). Electric
syntrophy through bDIET was first described by Summers et al. (2010) in defined
co-cultures of Geobacter metallireducens and Geobacter sulfurreducens with etha-
nol and fumarate as the sole electron donor and acceptor, respectively. G.
metallireducens, which can oxidize ethanol to obtain energy for growth but cannot
use fumarate as an electron acceptor, and G. sulfurreducens, which cannot use
ethanol as an electron donor but can reduce fumarate as an electron acceptor,
developed dense electrically conductive aggregates and syntrophically metabolize
ethanol with the reduction of fumarate in the co-cultures, even when an hyb-deleted
mutant strain of G. sulfurreducens, unable to utilize H2 or formate, was used.
However, deletion of the genes encoding OmcS (a multi-heme c-type cytochrome)
or PilA (the structural protein for e-pili) prevented the syntrophic metabolism
between the Geobacter strains and their growth. Studies with different mutant strains
and transcriptome assays proved the development of DIET via e-pili networks
associated with c-type cytochromes in the Geobacter co-cultures (Summers et al.
2010; Rotaru et al. 2012; Shrestha et al. 2013; Vargas et al. 2013).
Fig. 2.3 Direct interspecies electron transfer between syntrophic VFA-oxidizing bacteria and
hydrogenotrophic methanogens via biotic compounds (A) and abiotic conductive material (B)
methanogenesis of CO2 and that e-pili played an important role in building their
electrical connections. Although the genes for CO2 reduction were previously found
in Methanothrix species (Zhu et al. 2012), the electrotrophic conversion of CO2 to
CH4 by Methanothrix was first demonstrated by Rotaru et al. (2014b). This was a
32 C. Lee
Geobacter species to develop e-pili networks was likely an important factor that
facilitated the electro-syntrophic interactions in the DIET co-cultures (Martins et al.
2018). This corresponds to the fact that the existence of bDIET involved in
methanogenic systems was first observed in studies of conductive methanogenic
aggregates derived from brewery waste digesters (Morita et al. 2011; Rotaru et al.
2014b; Shrestha et al. 2014). Many (putative) exoelectrogenic bacteria other than
Geobacter species have been suggested (not yet proven) in recent studies to partic-
ipate in DIET with (putative) electrotrophic methanogens (Barua and Dhar 2017;
Park et al. 2018a). Therefore, in order to gain a comprehensive understanding of
DIET mechanisms and the involved microorganisms in diverse methanogenic envi-
ronments, more studies are needed, for example, in defined co-cultures or complex
mixed cultures with different electron donors other than ethanol.
Table 2.2 Electrical conductivity of abiotic conductive materials used for promoting DIET
Conductive materials Conductivity (mS/cm) Reference
Granular activated carbon 3–1200 Liu et al. (2012), Cheng et al. (2018)
Carbon cloth 4350 Lei et al. (2016)
Biochar 0.002–220 Chen et al. (2014b), Cheng et al. (2018)
Magnetite 105–106 Cornell and Schwertmann (2003)
Graphene 8.5 3 105 Lin et al. (2018)
Carbon nanotube >105 Yan et al. (2019)
Stainless steel 0.667 Li et al. (2017)
Polyaniline nanorod ~740 Hu et al. (2017)
G. Sulfurreducens e-pilia 51 Adhikari et al. (2016)
a
Pili from wild-type cells were measured in a hydrated state at pH 7
conductive materials that are large enough to serve as supporting media for attached
growth, while submicron/nanoparticles agglomerate together with microorganisms
to form aggregates such as flocs or granules (Baek et al. 2019).
Conductivity is the key characteristic of abiotic additives for promoting mDIET,
bypassing the need for electrical contacts via biological components necessary for
bDIET (Shrestha and Rotaru 2014). In support, magnetite nanoparticles were able to
compensate for the absence of OmcS, which, together with e-pili, forms electrical
conduits for DIET, in co-cultures of a mutant G. sulfurreducens strain deficient in
OmcS and G. metallireducens with ethanol as the electron donor (Liu et al. 2015). In
contrast to magnetite, which could only marginally compensate for the lack of PilA
(Liu et al. 2015), GAC was able to promote mDIET between the two Geobacter
species in co-cultures with ethanol, and to effectively compensate for the lack of both
PilA and OmcS (Liu et al. 2012). The findings of these studies, although not
performed with methanogenic cultures, confirm that abiotic conductive materials
can electrically connect DIET partners and stimulate their electric syntrophy, inde-
pendent of bDIET, and that the DIET mechanisms may differ among different
conductive additives with different characteristics, such as conductivity and size
(Baek et al. 2019). Biochar (Chen et al. 2014b) was shown to be comparably
effective to GAC (Liu et al. 2012) or carbon cloth (Chen et al. 2014a) in promoting
DIET, although the conductivity of biochar is approximately 1000-fold lower than
that of the others (Table 2.2). Furthermore, semiconductive hematite and conductive
magnetite showed similar methanogenesis-accelerating effects in DIET-stimulated
soil enrichment cultures with acetate or ethanol (Kato et al. 2012). By contrast, Baek
et al. (2015) and Zhuang et al. (2015) reported that the addition of iron oxides with
different conductivities (i.e., goethite, hematite, and magnetite) promoted DIET to
different extents in methanogenic cultures degrading whey and benzoate, respec-
tively. These results indicate that the DIET-promoting effect of a conductive additive
is not determined solely by its conductivity and will thus not continuously increase
in proportion to its electrical conductivity (Barua and Dhar 2017).
Following the studies in defined co-cultures, mostly pairing methanogens with
Geobacter species, which confirmed the existence of DIET in anaerobic
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 35
were noted in methanogenic reactors treating whey (Baek et al. 2015). The accu-
mulated evidence suggests that promoting mDIET by adding conductive materials
can provide a simple and effective means for enhancing methanogenic performance
and maintaining stable operation of AD processes.
Since the observation by Kato et al. (2012) that methanogenesis was accelerated by
adding (semi)conductive iron oxides in anaerobic mixed cultures with acetate or
ethanol, the potential of promoting mDIET as a means to improve the performance
of anaerobic digesters has gained much attention. As described in the preceding
section, many subsequent studies confirmed the widespread occurrence of DIET-
active methanogenic consortia in anaerobic environments and demonstrated the
DIET-promoting and methanogenesis-boosting effects of various conductive mate-
rials in defined co-cultures and mixed cultures. Adding inexpensive conductive
materials is currently considered a promising approach to enhance methanogenic
performance at a more fundamental level, and significant research efforts have
recently been made to investigate this possibility under more realistic conditions
for practical applications. This section describes the recent advances in engineering
DIET with conductive additives for enhanced methanogenesis in (semi-)continuous
anaerobic digesters treating complex organic wastes, including the latest studies at
pilot scale. Particular attention is given to the use of different (mostly carbon- or
iron-based) conductive additives and their effects on the digester performance.
GAC) with synthetic dairy wastewater in sequencing batch mode. They found that
GAC could accelerate the syntrophic methanogenesis of simple organics, such as
ethanol and VFAs, when efficient acidogenesis was achieved in the presence of
magnetite.
Zhang et al. (2020c) and Guo et al. (2020a) reported significant increases in
chemical oxygen demand (COD) removal and methane production in upflow anaer-
obic sludge blanket (UASB) reactors treating low-strength synthetic wastewater
containing glucose (500 mg COD/L) at psychrophilic and mesophilic temperatures,
respectively. Xu et al. (2015) compared the effects of GAC (0.84–2.00 mm) and
powdered activated carbon (PAC; 75–177 μm) on the anaerobic treatment of
synthetic brewery wastewater in UASB reactors. Both GAC and PAC enhanced
the methanogenic degradation of ethanol and VFAs and facilitated the enrichment of
methanogens, while PAC with more abundant micro-mesoporous structures for the
colonization of DIET partners induced a greater increase in methane production,
particularly at higher organic loading rates (OLRs). Mei et al. (2018) reported, in a
study of the DIET-stimulated methanogenic degradation of synthetic soft drink
processing wastewater using upflow packed bed reactors (UPBRs), that the daily
methane production relative to the biomass concentration was 43.3% and 31.5%
greater in the digesters filled with GAC and anthracite, respectively, than in the
control digester with non-conductive ceramic media. They suggested a syntrophic
association between a novel Geobacter population and Methanothrix in the DIET-
stimulated digesters based on the metagenomic analysis results.
Additionally, several studies have reported the promotion of mDIET with acti-
vated carbon in (semi-)continuous digesters treating real waste/wastewater of dif-
ferent characteristics. Dang et al. (2016) compared four carbon-based conductive
materials (GAC, carbon cloth, carbon felt, and graphite rod; with similar geometric
surface areas of ~1000 cm2/L) for their effects on the performance of anaerobic
sequencing batch reactors (SBRs) treating commercial dog food (as a proxy for food
waste). The digesters supplemented with GAC, carbon cloth, and carbon felt showed
better and more stable AD performance than the control digesters with polyester
cloth or no material added, particularly at higher OLRs, whereas adding graphite
rods was ineffective in promoting or stabilizing methanogenesis.
Usman et al. (2019) and Yang et al. (2020b) studied the effect of adding GAC in
anaerobic digesters treating hydrothermal liquefaction wastewaters from sewage
sludge (SBR) and swine manure (UPBR), respectively. Both studies claimed that
GAC enhanced the methanogenic performance of the digesters by promoting DIET,
although the adsorption of inhibitors by GAC also must have contributed to the
improved methane yield. Lei et al. (2019) reported that GAC dosing promoted DIET
involving electroactive Geobacter and Methanosarcina and enhanced the
methanogenic performance in UASB reactors fed with municipal solid waste incin-
eration leachate. The leachate was effectively treated without a lag period in the
GAC-added digester at increasing OLRs from 5.9 to 36.7 g COD/Ld, while the
control digester without GAC failed within the first 17 days of operation at the
lowest OLR of 5.9 g COD/Ld. Zhang et al. (2020b) compared low-temperature
(16.5 2.0 C) UASB reactors with and without GAC addition for the
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 41
2.3.1.2 Biochar
Carbon cloth, felt, and fiber with high electrical conductivity have also been dem-
onstrated to promote syntrophic metabolism and accelerate methanogenesis in
several (semi-)continuous digesters fed with synthetic wastewaters with different
organic compounds as the main electron donor, such as ethanol (Zhao et al. 2015),
butanol (Zhao et al. 2017c), glucose (Xu et al. 2016), a propionate/butyrate mixture
(Barua et al. 2018), and a glucose/glycine mixture (Feng et al. 2020). As mentioned
above, carbon cloth and felt were tested, together with other conductive additives, in
a study using SBRs treating commercial dog food, and their effectiveness in
promoting DIET and improving digester performance and stability was demon-
strated (Dang et al. 2016). Lei et al. (2016) reported in a study using UASB reactors
treating municipal solid waste incineration leachate, that the carbon cloth-added
digester maintained stable performance with high organic removal at OLRs up to
49.4 g COD/Ld, where the control digester without carbon cloth failed with an
accumulation of VFAs. They found that the addition of carbon cloth enriched
potential exoelectrogenic bacteria and electrotrophic methanogens and increased
sludge conductivity by approximately two-fold, indicating the development of a
DIET-active methanogenic community.
Ambuchi et al. (2017) assessed the effects of adding multi-well carbon nanotubes or
iron oxide (Fe2O3) nanoparticles in expanded granular sludge bed reactors treating
beet sugar wastewater. The digester supplemented with Fe2O3 nanoparticles (0.75-g/
L dose) produced more methane than the one with carbon nanotubes (1.5-g/L dose),
while both digesters showed significantly higher methane production than the
control without conductive additive, although the cumulative methane content in
biogas was as low as 32–33% in all three digesters. Both carbon nanotubes and
Fe2O3 nanoparticles were suggested to provide electrical connections between DIET
partners and accelerate methane production. However, the former promoted the
enrichment of bacteria rather than methanogens, whereas the latter showed the
opposite tendency. Notably, carbon nanotubes have been reported to enhance
methanogenesis in pure cultures of methanogens, independently of DIET (Salvador
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 43
et al. 2017). This effect was attributable to the reduction of the oxidation-reduction
potential (only in the presence of reducing agents) and provision of large
electroactive surface area for microbial adhesion, both of which are beneficial for
the growth of methanogens, by the addition of carbon nanotubes (Martins et al.
2018). The market price of carbon nanotubes is US$100–150/kg (Fan et al. 2020),
which is 10–100 times more expensive than activated carbon (Yunus et al. 2020).
Hence, its continuous addition to digesters, despite its demonstrated effectiveness, is
not economically feasible.
2.3.1.5 Graphite
Jin et al. (2019) tested the effect of magnetite addition in UASB reactors treating
sulfate-rich synthetic wastewater containing glucose while varying the COD/sulfate
ratio from 5.0 to 2.5. They observed simultaneous enhancement of methanogenesis
and sulfate removal in the magnetite-supplemented digester and suggested the
establishment of electric syntrophy via DIET between Fe(III)/sulfate-reducing bac-
teria and methanogens. By contrast, Liu et al. (2019) reported that the sulfate-
reducing activity was reduced by the addition of magnetite, while VFA degradation
and methanogenesis were promoted, in anaerobic SBRs fed with sulfate-rich syn-
thetic wastewater (COD/sulfate ratio ¼ 6.0) containing ethanol, acetate, and
propionate.
Furthermore, a more recent study that examined the effect of magnetite in
anaerobic CSTRs treating a sulfur-rich organic waste mixture (Ulva biomass and
cheese whey; 0.8% w/w dry weight) observed no apparent enhancement of
methanogenic performance by the addition of magnetite (Jung et al. 2020). It was
found instead that the H2S content in biogas decreased remarkably with magnetite
addition through anaerobic oxidation of sulfide to elemental sulfur, which was
suggested to be coupled via DIET with electrotrophic methanogenesis. These
different observations may be attributed in part to the differences in experimental
conditions, such as inoculum source, magnetite dose, substrate/electron donor, and
reactor type/operation mode, and at the same time underscore the need for more
studies in sulfate/sulfur-rich environments, where sulfate-reducing bacteria can
compete strongly with methanogens for common electron donors.
The DIET-promoting effect of magnetite was also demonstrated with different real
waste streams in (semi-)continuous anerobic digesters. Baek et al. (2016) demon-
strated the beneficial effect of magnetite on the methanogenic performance and
stability of anaerobic CSTRs treating whey in long-term experiments over a
one-year period. As an extension study, they investigated the potential of magnetic
separation and recycling of magnetite from the digester effluent as a means to
maintain enhanced stable digester performance without continuous supplementation
with magnetite (Baek et al. 2017). Magnetite recycling effectively maintained
46 C. Lee
promoted DIET and methanogenic activities over a long period of time (>250 days)
and helped to maintain high biomass retention by returning active microorganisms
associated with magnetite to the digester. In both of the above studies, magnetite
addition significantly affected the microbial community structure, and Methanothrix
was suggested to be responsible for electrotrophic methanogenesis via DIET.
Lei et al. (2018) reported the enhancement of methane production and COD
removal with the addition of magnetite in UASB reactors treating municipal solid
waste incineration leachate. The microbial community structure was significantly
different between the digesters with and without magnetite addition, and
exoelectrogenic bacteria and electrotrophic methanogens (Methanosarcina and
Methanothrix), which were potentially involved in DIET, were enriched in the
magnetite-supplemented digester. Wang et al. (2018b) compared anaerobic SBRs
added with different doses of magnetite (0–0.3 g/L) for the treatment of Fischer-
Tropsch wastewater from a coal-to-liquids plant. The digesters supplemented with
0.1 or 0.2 g/L magnetite showed a significantly higher organic removal and methane
production compared to the control digester, with 0.2 g/L being the optimal magne-
tite dose, while the digester dosed with 0.3 g/L magnetite showed the worst
performance. Magnetite addition promoted the enrichment of Geobacter and
Methanothrix and their electric syntrophy via DIET, particularly at the optimal
magnetite dose.
New and other conductive materials besides the above-mentioned carbon- and iron-
based materials have been evaluated for effectiveness in promoting DIET during AD
in recent studies. However, most of the studies have been performed in batch mode
with synthetic wastewater, and limited information is available on the practicability
of the materials in (semi-)continuous processes. Polyaniline, a conductive polymer,
was applied in the form of a hydrogel (Zhou et al. 2021) and nanorod (Hu et al. 2017)
to promote DIET in the methanogenic treatment of synthetic wastewater containing
sucrose in batch cultures, and proved effective in accelerating methane production.
Wang et al. (2021b) prepared magnetite-contained biochar using iron-rich Fenton
sludge and demonstrated its promoting effect on DIET and methanogenesis in batch
cultures anaerobically treating synthetic dairy wastewater.
Yang et al. (2020a) reported that composite GAC-MnO2 was significantly more
effective than GAC in enhancing electron transfer between syntrophic microorgan-
isms and thus methanogenesis during the batch AD of synthetic wastewater
containing starch. Different non-iron transition metal compounds (i.e., WO2, WC,
W2N, W18O49, TaOx, Nb2O5, and HfO2) were shown to be effective in stimulating
DIET and enhancing methane production in anaerobic batch cultures fed with dairy
manure (Wang et al. 2020c; Yun et al. 2021). Guo et al. (2020b) examined the effect
of nickel foam addition on the batch methanation of synthetic wastewater containing
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 47
ethanol and reported that nickel foam effectively promoted DIET and accelerated
methane production.
Liu et al. (2020) compared the performance of three anaerobic integrated floating
fixed-film activated sludge (An-IFFAS) reactors filled with high-density polyethyl-
ene (HDPE)-based carriers with 3%, 5%, and 7% (w/w) graphite contents with two
control reactors, one anaerobic floc reactor and one An-IFFAS reactor filled with
plain HDPE carriers, for the continuous anaerobic treatment of synthetic wastewater
containing ethanol. All graphite-modified carrier-filled reactors performed signifi-
cantly better than the controls, and the reactors filled with the carriers with higher
graphite contents showed greater methanogenic performance. The addition of
graphite-modified carriers enriched potential DIET partners, Geobacter and
Methanothrix, which could accelerate syntrophic methanogenesis.
Several studies applied the addition of conductive materials combined with the
application of external voltage as a strategy to further promote DIET and boost
syntrophic methanogenesis (Baek et al. 2020, 2021; Vu et al. 2020). Combining AD
with a microbial electrolysis cell (AD-MEC) has been suggested as a way to improve
AD performance, because the cathodic CO2 conversion to CH4 by electrotrophic or
hydrogenotrophic methanogens increases methane production. However, it is yet to
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 49
that have been used to confirm the occurrence of DIET in methanogenic systems and
the need for the development of suitable DIET characterization methods in detail
(Van Steendam et al. 2019). Technologies for engineering DIET for enhanced AD
are still in the embryonic stage. Quantitative methods to monitor DIET activity will
not only help to advance our fundamental understanding of DIET but also to enable
better engineering and operation of DIET-aided AD processes. Furthermore, such
methods will provide the ability to unravel the conflicting results in the literature.
Although conductive materials have generally been reported to promote DIET
and enhance the methanogenic conversion of organic matter, a few studies have
demonstrated no or rather inhibitory effects for some carbon- and metal-based
conductive materials, such as ferrihydrite, carbon black, magnesium oxide, and
silver nanoparticles (Martins et al. 2018; Baek et al. 2019). It is understandable
that conductive materials may exert different effects on DIET and thus
methanogenesis according to their different physicochemical properties (e.g., elec-
tric conductivity, crystallinity, solubility, surface area, porosity, morphology, and
point of zero charge) and the digester environmental conditions (e.g., substrate,
inoculum, oxidation-reduction potential, pH, temperature, inhibitory substances,
and other operational parameters). However, the underlying reasons for the
conflicting results, especially for the inhibitory effects, are not well understood,
although their understanding is vital for effectively engineering DIET for enhanced
methanogenesis.
Another point that needs to be addressed is the effect of the addition of conductive
materials on electro-syntrophic interactions other than electrotrophic
methanogenesis. Many studies have reported enhancement of microbial and enzy-
matic activities related to hydrolysis, fermentation, and acetoclastic methanogenesis,
besides those associated with CO2-reducing methanogenesis, in methanogenic cul-
tures supplemented with conductive additives. These observations are not surprising
given that highly diverse microorganisms, including many (potential) electroactive
microorganisms, coexist and interact in mixed-culture methanogenic systems. It is
not unlikely that mDIET-promoting approaches may affect other microbial redox
processes, either directly or indirectly.
For example, anoxygenic photosynthesis (green sulfur bacteria Prosthecochloris)
can be coupled to anaerobic respiration (Geobacter) via DIET under anoxic condi-
tions (Ha et al. 2017). This syntrophic anaerobic photosynthesis may occur in
methanogenic environments, although its effect on the overall AD process would
be limited, particularly in field-scale AD plants run under completely dark conditions
with no light for photosynthesis. Furthermore, a recent study reported that magnetite
addition did not enhance methane production but greatly reduced the H2S content in
biogas during the semi-continuous AD of sulfur-rich organic waste (Jung et al.
2020). The authors found that H2S was removed by microbial oxidation to S0 and
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 51
Acknowledgement This work was supported by a grant from the National Research Foundation of
Korea (NRF-2020R1A2C2004368).
References
Adhikari RY, Malvankar NS, Tuominen MT, Lovley DR (2016) Conductivity of individual
Geobacter pili. RSC Adv 6:8354–8357
Ambuchi JJ, Zhang Z, Shan L, Liang D, Zhang P, Feng Y (2017) Response of anaerobic granular
sludge to iron oxide nanoparticles and multi-wall carbon nanotubes during beet sugar industrial
wastewater treatment. Water Res 117:87–94
Baek G, Kim J, Cho K, Bae H, Lee C (2015) The biostimulation of anaerobic digestion with
(semi)conductive ferric oxides: their potential for enhanced biomethanation. Appl Microbiol
Biotechnol 99:10355–10366
Baek G, Kim J, Lee C (2016) A long-term study on the effect of magnetite supplementation in
continuous anaerobic digestion of dairy effluent – enhancement in process performance and
stability. Bioresour Technol 222:344–354
Baek G, Jung H, Kim J, Lee C (2017) A long-term study on the effect of magnetite supplementation
in continuous anaerobic digestion of dairy effluent – magnetic separation and recycling of
magnetite. Bioresour Technol 241:830–840
Baek G, Kim J, Kim J, Lee C (2018) Role and potential of direct interspecies electron transfer in
anaerobic digestion. Energies 11:107
Baek G, Kim J, Lee C (2019) A review of the effects of iron compounds on methanogenesis in
anaerobic environments. Renew Sust Energ Rev 113:109282
Baek G, Kim J, Kim J, Lee C (2020) Individual and combined effects of magnetite addition and
external voltage application on anaerobic digestion of dairy wastewater. Bioresour Technol 297:
122443
Baek G, Saikaly PE, Logan BE (2021) Addition of a carbon fiber brush improves anaerobic
digestion compared to external voltage application. Water Res 188:116575
Barua S, Dhar BR (2017) Advances towards understanding and engineering direct interspecies
electron transfer in anaerobic digestion. Bioresour Technol 244:698–707
Barua S, Zakaria BS, Dhar BR (2018) Enhanced methanogenic co-degradation of propionate and
butyrate by anaerobic microbiome enriched on conductive carbon fibers. Bioresour Technol
266:259–266
Boone DR, Castenholz RW (eds) (2001) Bergey's manual of systematic bacteriology: the archaea
and the deeply branching and phototrophic bacteria, vol 1, 2nd edn. Springer, New York
52 C. Lee
Boone DR, Johnson RL, Liu Y (1989) Diffusion of the interspecies electron carriers H2 and formate
in methanogenic ecosystems and its implications in the measurement of km for H2 or formate
uptake. Appl Environ Microbiol 55:1735
Capson-Tojo G, Moscoviz R, Ruiz D, Santa-Catalina G, Trably E, Rouez M, Crest M, Steyer J-P,
Bernet N, Delgenès J-P, Escudié R (2018) Addition of granular activated carbon and trace
elements to favor volatile fatty acid consumption during anaerobic digestion of food waste.
Bioresour Technol 260:157–168
Chen S, Rotaru A-E, Liu F, Philips J, Woodard TL, Nevin KP, Lovley DR (2014a) Carbon cloth
stimulates direct interspecies electron transfer in syntrophic co-cultures. Bioresour Technol 173:
82–86
Chen S, Rotaru A-E, Shrestha PM, Malvankar NS, Liu F, Fan W, Nevin KP, Lovley DR (2014b)
Promoting interspecies electron transfer with biochar. Sci Rep 4:5019
Chen Q, Liu C, Liu X, Sun D, Li P, Qiu B, Dang Y, Karpinski NA, Smith JA, Holmes DE (2020)
Magnetite enhances anaerobic digestion of high salinity organic wastewater. Environ Res 189:
109884
Cheng Q, Francis L, Call DF (2018) Amending anaerobic bioreactors with pyrogenic carbonaceous
materials: the influence of material properties on methane generation. Environ Sci: Water Res
Technol 4:1794–1806
Chowdhury B, Lin L, Dhar BR, Islam MN, McCartney D, Kumar A (2019) Enhanced biomethane
recovery from fat, oil, and grease through co-digestion with food waste and addition of
conductive materials. Chemosphere 236:124362
Cornell RM, Schwertmann U (2003) The iron oxides: structure, properties, reactions, occurrences
and uses. Wiley, Weinheim
Cruz Viggi C, Rossetti S, Fazi S, Paiano P, Majone M, Aulenta F (2014) Magnetite particles
triggering a faster and more robust syntrophic pathway of methanogenic propionate degrada-
tion. Environ Sci Technol 48:7536–7543
Dang Y, Holmes DE, Zhao Z, Woodard TL, Zhang Y, Sun D, Wang L-Y, Nevin KP, Lovley DR
(2016) Enhancing anaerobic digestion of complex organic waste with carbon-based conductive
materials. Bioresour Technol 220:516–522
Dang Y, Sun D, Woodard TL, Wang L-Y, Nevin KP, Holmes DE (2017) Stimulation of the
anaerobic digestion of the dry organic fraction of municipal solid waste (OFMSW) with carbon-
based conductive materials. Bioresour Technol 238:30–38
de Bok FAM, Plugge CM, Stams AJM (2004) Interspecies electron transfer in methanogenic
propionate degrading consortia. Water Res 38:1368–1375
Dubé C-D, Guiot SR (2015) Direct interspecies electron transfer in anaerobic digestion: a
review. In: Gubitz G, Bauer A, Bochmann G, Gronauer A, Weiss S (eds) Biogas science and
technology. Springer, Cham, pp 101–115
Enzmann F, Mayer F, Rother M, Holtmann D (2018) Methanogens: biochemical background and
biotechnological applications. AMB Express 8:1
Evans PN, Parks DH, Chadwick GL, Robbins SJ, Orphan VJ, Golding SD, Tyson GW (2015)
Methane metabolism in the archaeal phylum bathyarchaeota revealed by genome-centric
metagenomics. Science 350:434
Fan J, Li G, Deng S, Deng C, Wang Z, Zhang Z (2020) Effect of carbon nanotube and styrene-
acrylic emulsion additives on microstructure and mechanical characteristics of cement paste.
Materials 13:2807
Felchner-Zwirello M, Winter J, Gallert C (2013) Interspecies distances between propionic acid
degraders and methanogens in syntrophic consortia for optimal hydrogen transfer. Appl
Microbiol Biotechnol 97:9193–9205
Feng D, Xia A, Liao Q, Nizami A-S, Sun C, Huang Y, Zhu X, Zhu X (2020) Carbon cloth facilitates
semi-continuous anaerobic digestion of organic wastewater rich in volatile fatty acids from dark
fermentation. Environ Pollut 272:116030
Fernandez AS, Hashsham SA, Dollhopf SL, Raskin L, Glagoleva O, Dazzo FB, Hickey RF, Criddle
CS, Tiedje JM (2000) Flexible community structure correlates with stable community function
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 53
Lü C, Shen Y, Li C, Zhu N, Yuan H (2020) Redox-active biochar and conductive graphite stimulate
methanogenic metabolism in anaerobic digestion of waste-activated sludge: beyond direct
interspecies electron transfer. ACS Sustain Chem Eng 8:12626–12636
Lubitz W, Ogata H, Rüdiger O, Reijerse E (2014) Hydrogenases. Chem Rev 114:4081–4148
Luo C, Lü F, Shao L, He P (2015) Application of eco-compatible biochar in anaerobic digestion to
relieve acid stress and promote the selective colonization of functional microbes. Water Res 68:
710–718
Martins G, Salvador AF, Pereira L, Alves MM (2018) Methane production and conductive
materials: a critical review. Environ Sci Technol 52:10241–10253
Mei R, Nobu MK, Narihiro T, Yu J, Sathyagal A, Willman E, Liu W-T (2018) Novel Geobacter
species and diverse methanogens contribute to enhanced methane production in media-added
methanogenic reactors. Water Res 147:403–412
Morita M, Malvankar NS, Franks AE, Summers ZM, Giloteaux L, Rotaru AE, Rotaru C, Lovley
DR (2011) Potential for direct interspecies electron transfer in methanogenic wastewater
digester aggregates. MBio 2:e00159–e00111
Nielsen CF, Lange L, Meyer AS (2019) Classification and enzyme kinetics of formate dehydroge-
nases for biomanufacturing via CO2 utilization. Biotechnol Adv 37:107408
Park J-H, Kang H-J, Park K-H, Park H-D (2018a) Direct interspecies electron transfer via conduc-
tive materials: a perspective for anaerobic digestion applications. Bioresour Technol 254:300–
311
Park J-H, Park J-H, Je Seong H, Sul WJ, Jin K-H, Park H-D (2018b) Metagenomic insight into
methanogenic reactors promoting direct interspecies electron transfer via granular activated
carbon. Bioresour Technol 259:414–422
Petrovic S (2021) Overpotential. Electrochemistry crash course for engineers. Springer, Cham, pp
59–64
Prüsse U, Hähnlein M, Daum J, Vorlop K-D (2000) Improving the catalytic nitrate reduction. Catal
Today 55:79–90
Rotaru A-E, Shrestha PM, Liu F, Ueki T, Nevin K, Summers ZM, Lovley DR (2012) Interspecies
electron transfer via hydrogen and formate rather than direct electrical connections in cocultures
of Pelobacter carbinolicus and Geobacter sulfurreducens. Appl Environ Microbiol 78:7645–
7651
Rotaru A-E, Shrestha PM, Liu F, Markovaite B, Chen S, Nevin KP, Lovley DR (2014a) Direct
interspecies electron transfer between Geobacter metallireducens and Methanosarcina barkeri.
Appl Environ Microbiol 80:4599–4605
Rotaru A-E, Shrestha PM, Liu F, Shrestha M, Shrestha D, Embree M, Zengler K, Wardman C,
Nevin KP, Lovley DR (2014b) A new model for electron flow during anaerobic digestion: direct
interspecies electron transfer to methanosaeta for the reduction of carbon dioxide to methane.
Energy Environ Sci 7:408–415
Salvador AF, Martins G, Melle-Franco M, Serpa R, Stams AJ, Cavaleiro AJ, Pereira MA, Alves
MM (2017) Carbon nanotubes accelerate methane production in pure cultures of methanogens
and in a syntrophic coculture. Environ Microbiol 19:2727–2739
Schink B (1997) Energetics of syntrophic cooperation in methanogenic degradation. Microbiol Mol
Biol Rev 61:262–280
Schink B, Montag D, Keller A, Müller N (2017) Hydrogen or formate: alternative key players in
methanogenic degradation. Environ Microbiol Rep 9:189–202
Schmidt JE, Ahring BK (1993) Effects of hydrogen and formate on the degradation of propionate
and butyrate in thermophilic granules from an upflow anaerobic sludge blanket reactor. Appl
Environ Microbiol 59:2546
Schmidt JE, Ahring BK (1995) Interspecies electron transfer during propionate and butyrate
degradation in mesophilic, granular sludge. Appl Environ Microbiol 61:2765
Shrestha PM, Rotaru A-E (2014) Plugging in or going wireless: strategies for interspecies electron
transfer. Front Microbiol 5:237
56 C. Lee
Shrestha PM, Rotaru A-E, Summers ZM, Shrestha M, Liu F, Lovley DR (2013) Transcriptomic and
genetic analysis of direct interspecies electron transfer. Appl Environ Microbiol 79:2397–2404
Shrestha PM, Malvankar NS, Werner JJ, Franks AE, Elena-Rotaru A, Shrestha M, Liu F, Nevin KP,
Angenent LT, Lovley DR (2014) Correlation between microbial community and granule
conductivity in anaerobic bioreactors for brewery wastewater treatment. Bioresour Technol
174:306–310
Smith KS, Ingram-Smith C (2007) Methanosaeta, the forgotten methanogen? Trends Microbiol 15:
150–155
Stams AJM, Plugge CM (2009) Electron transfer in syntrophic communities of anaerobic bacteria
and archaea. Nat Rev Microbiol 7:568–577
Stams AJM, Bok FAM, Plugge CM, Eekert MHA, Dolfing J, Schraa G (2006) Exocellular electron
transfer in anaerobic microbial communities. Environ Microbiol 8:371–382
Storck T, Virdis B, Batstone DJ (2016) Modelling extracellular limitations for mediated versus
direct interspecies electron transfer. ISME J 10:621–631
Stroot PG, McMahon KD, Mackie RI, Raskin L (2001) Anaerobic codigestion of municipal solid
waste and biosolids under various mixing conditions – i. digester performance. Water Res 35:
1804–1816
Summers ZM, Fogarty HE, Leang C, Franks AE, Malvankar NS, Lovley DR (2010) Direct
exchange of electrons within aggregates of an evolved syntrophic coculture of anaerobic
bacteria. Science 330:1413–1415
Sun T, Levin BDA, Guzman JJL, Enders A, Muller DA, Angenent LT, Lehmann J (2017) Rapid
electron transfer by the carbon matrix in natural pyrogenic carbon. Nature Comm 8:14873
Thiele JH, Zeikus JG (1988) Control of interspecies electron flow during anaerobic digestion:
significance of formate transfer versus hydrogen transfer during syntrophic methanogenesis in
flocs. Appl Environ Microbiol 54:20
Tian T, Qiao S, Li X, Zhang M, Zhou J (2017) Nano-graphene induced positive effects on
methanogenesis in anaerobic digestion. Bioresour Technol 224:41–47
Usman M, Hao S, Chen H, Ren S, Tsang DCW, O-Thong S, Luo G, Zhang S (2019) Molecular and
microbial insights towards understanding the anaerobic digestion of the wastewater from
hydrothermal liquefaction of sewage sludge facilitated by granular activated carbon (GAC).
Environ Int 133:105257
Van Steendam C, Smets I, Skerlos S, Raskin L (2019) Improving anaerobic digestion via direct
interspecies electron transfer requires development of suitable characterization methods. Curr
Opin Biotech 57:183–190
Vanwonterghem I, Evans PN, Parks DH, Jensen PD, Woodcroft BJ, Hugenholtz P, Tyson GW
(2016) Methylotrophic methanogenesis discovered in the archaeal phylum Verstraetearchaeota.
Nature Microbiol 1:16170
Vargas M, Malvankar NS, Tremblay P-L, Leang C, Smith JA, Patel P, Snoeyenbos-West O, Nevin
KP, Lovley DR (2013) Aromatic amino acids required for pili conductivity and long-range
extracellular electron transport in Geobacter sulfurreducens. MBio 4:e00105–e00113
Vu MT, Noori MT, Min B (2020) Conductive magnetite nanoparticles trigger syntrophic methane
production in single chamber microbial electrochemical systems. Bioresour Technol 296:
122265
Wang C, Liu Y, Gao X, Chen H, Xu X, Zhu L (2018a) Role of biochar in the granulation of
anaerobic sludge and improvement of electron transfer characteristics. Bioresour Technol 268:
28–35
Wang D, Han Y, Han H, Li K, Xu C, Zhuang H (2018b) New insights into enhanced anaerobic
degradation of Fischer-Tropsch wastewater with the assistance of magnetite. Bioresour Technol
257:147–156
Wang T, Zhang D, Dai L, Dong B, Dai X (2018c) Magnetite triggering enhanced direct interspecies
electron transfer: a scavenger for the blockage of electron transfer in anaerobic digestion of
high-solids sewage sludge. Environ Sci Technol 52:7160–7169
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 57
Wang C, Liu Y, Jin S, Chen H, Xu X, Wang Z, Xing B, Zhu L (2019a) Responsiveness extracellular
electron transfer (EET) enhancement of anaerobic digestion system during start-up and starva-
tion recovery stages via magnetite addition. Bioresour Technol 272:162–170
Wang Z, Yun S, Xu H, Wang C, Zhang Y, Chen J, Jia B (2019b) Mesophilic anaerobic co-digestion
of acorn slag waste with dairy manure in a batch digester: focusing on mixing ratios and
bio-based carbon accelerants. Bioresour Technol 286:121394
Wang C, Wang C, Liu J, Han Z, Xu Q, Xu X, Zhu L (2020a) Role of magnetite in methanogenic
degradation of different substances. Bioresour Technol 314:123720
Wang G, Gao X, Li Q, Zhao H, Liu Y, Wang XC, Chen R (2020b) Redox-based electron exchange
capacity of biowaste-derived biochar accelerates syntrophic phenol oxidation for
methanogenesis via direct interspecies electron transfer. J Hazard Mater 390:121726
Wang Z, Yun S, Shi J, Han F, Liu B, Wang R, Li X (2020c) Critical evidence for direct interspecies
electron transfer with tungsten-based accelerants: an experimental and theoretical investigation.
Bioresour Technol 311:123519
Wang C, Liu Y, Wang C, Xing B, Zhu S, Huang J, Xu X, Zhu L (2021a) Biochar facilitates rapid
restoration of methanogenesis by enhancing direct interspecies electron transfer after high
organic loading shock. Bioresour Technol 320:124360
Wang M, Zhao Z, Zhang Y (2021b) Magnetite-contained biochar derived from Fenton sludge
modulated electron transfer of microorganisms in anaerobic digestion. J Hazard Mater 403:
123972
Xu S, He C, Luo L, Lü F, He P, Cui L (2015) Comparing activated carbon of different particle sizes
on enhancing methane generation in upflow anaerobic digester. Bioresour Technol 196:606–
612
Xu H, Wang C, Yan K, Wu J, Zuo J, Wang K (2016) Anaerobic granule-based biofilms formation
reduces propionate accumulation under high h2 partial pressure using conductive carbon felt
particles. Bioresour Technol 216:677–683
Xu S, Zhang W, Zuo L, Qiao Z, He P (2020) Comparative facilitation of activated carbon and
goethite on methanogenesis from volatile fatty acids. Bioresour Technol 302:122801
Yamada C, Kato S, Ueno Y, Ishii M, Igarashi Y (2015) Conductive iron oxides accelerate
thermophilic methanogenesis from acetate and propionate. J Biosci Bioeng 119:678–682
Yan W, Shen N, Xiao Y, Chen Y, Sun F, Kumar Tyagi V, Zhou Y (2017) The role of conductive
materials in the start-up period of thermophilic anaerobic system. Bioresour Technol 239:336–
344
Yan W, Lu D, Liu J, Zhou Y (2019) The interactive effects of ammonia and carbon nanotube on
anaerobic digestion. Chem Eng J 372:332–340
Yang Z, Guo R, Shi X, Wang C, Wang L, Dai M (2016) Magnetite nanoparticles enable a rapid
conversion of volatile fatty acids to methane. RSC Adv 6:25662–25668
Yang Y, Zhang Y, Li Z, Zhao Z, Quan X, Zhao Z (2017) Adding granular activated carbon into
anaerobic sludge digestion to promote methane production and sludge decomposition. J Clean
Prod 149:1101–1108
Yang B, Xu H, Liu Y, Li F, Song X, Wang Z, Sand W (2020a) Role of GAC-MnO2 catalyst for
triggering the extracellular electron transfer and boosting CH4 production in syntrophic
methanogenesis. Chem Eng J 383:123211
Yang L, Si B, Zhang Y, Watson J, Stablein M, Chen J, Zhang Y, Zhou X, Chu H (2020b)
Continuous treatment of hydrothermal liquefaction wastewater in an anaerobic biofilm reactor:
potential role of granular activated carbon. J Clean Prod 276:122836
Yee MO, Rotaru A-E (2020) Extracellular electron uptake in methanosarcinales is independent of
multiheme c-type cytochromes. Sci Rep 10:372
Yee MO, Snoeyenbos-West OL, Thamdrup B, Ottosen LDM, Rotaru A-E (2019) Extracellular
electron uptake by two Methanosarcina species. Front Energy Res 7:29
Yin Q, He K, Liu A, Wu G (2017a) Enhanced system performance by dosing ferroferric oxide
during the anaerobic treatment of tryptone-based high-strength wastewater. Appl Microbiol
Biotechnol 101:3929–3939
58 C. Lee
Yin Q, Miao J, Li B, Wu G (2017b) Enhancing electron transfer by ferroferric oxide during the
anaerobic treatment of synthetic wastewater with mixed organic carbon. Int Biodeterior Bio-
degradation 119:104–110
Yin Q, Yang S, Wang Z, Xing L, Wu G (2018) Clarifying electron transfer and metagenomic
analysis of microbial community in the methane production process with the addition of
ferroferric oxide. Chem Eng J 333:216–225
Yun S, Xing T, Han F, Shi J, Wang Z, Fan Q, Xu H (2021) Enhanced direct interspecies electron
transfer with transition metal oxide accelerants in anaerobic digestion. Bioresour Technol 320:
124294
Yunus ZM, Al-Gheethi A, Othman N, Hamdan R, Ruslan NN (2020) Removal of heavy metals
from mining effluents in tile and electroplating industries using honeydew peel activated carbon:
a microstructure and techno-economic analysis. J Clean Prod 251:119738
Zhang J, Lu Y (2016) Conductive Fe3O4 nanoparticles accelerate syntrophic methane production
from butyrate oxidation in two different lake sediments. Front Microbiol 7:1316
Zhang J, Mao L, Zhang L, Loh K-C, Dai Y, Tong YW (2017) Metagenomic insight into the
microbial networks and metabolic mechanism in anaerobic digesters for food waste by incor-
porating activated carbon. Sci Rep 7:11293
Zhang L, Zhang J, Loh K-C (2018) Activated carbon enhanced anaerobic digestion of food waste –
laboratory-scale and pilot-scale operation. Waste Manag 75:270–279
Zhang J, Lu T, Wang Z, Wang Y, Zhong H, Shen P, Wei Y (2019a) Effects of magnetite on
anaerobic digestion of swine manure: attention to methane production and fate of antibiotic
resistance genes. Bioresour Technol 291:121847
Zhang M, Ma Y, Ji D, Li X, Zhang J, Zang L (2019b) Synergetic promotion of direct interspecies
electron transfer for syntrophic metabolism of propionate and butyrate with graphite felt in
anaerobic digestion. Bioresour Technol 287:121373
Zhang J, Zhang R, Wang H, Yang K (2020a) Direct interspecies electron transfer stimulated by
granular activated carbon enhances anaerobic methanation efficiency from typical kitchen waste
lipid-rapeseed oil. Sci Total Environ 704:135282
Zhang Y, Guo B, Zhang L, Liu Y (2020b) Key syntrophic partnerships identified in a granular
activated carbon amended uasb treating municipal sewage under low temperature conditions.
Bioresour Technol 312:123556
Zhang Y, Zhang L, Guo B, Zhou Y, Gao M, Sharaf A, Liu Y (2020c) Granular activated carbon
stimulated microbial physiological changes for enhanced anaerobic digestion of municipal
sewage. Chem Eng J 400:125838
Zhao Z, Zhang Y (2019) Application of ethanol-type fermentation in establishment of direct
interspecies electron transfer: a practical engineering case study. Renew Energ 136:846–855
Zhao Z, Zhang Y, Woodard TL, Nevin KP, Lovley DR (2015) Enhancing syntrophic metabolism in
up-flow anaerobic sludge blanket reactors with conductive carbon materials. Bioresour Technol
191:140–145
Zhao Z, Zhang Y, Holmes DE, Dang Y, Woodard TL, Nevin KP, Lovley DR (2016a) Potential
enhancement of direct interspecies electron transfer for syntrophic metabolism of propionate
and butyrate with biochar in up-flow anaerobic sludge blanket reactors. Bioresour Technol 209:
148–156
Zhao Z, Zhang Y, Yu Q, Dang Y, Li Y, Quan X (2016b) Communities stimulated with ethanol to
perform direct interspecies electron transfer for syntrophic metabolism of propionate and
butyrate. Water Res 102:475–484
Zhao Z, Li Y, Quan X, Zhang Y (2017a) New application of ethanol-type fermentation: stimulating
methanogenic communities with ethanol to perform direct interspecies electron transfer. ACS
Sustain Chem Eng 5:9441–9453
Zhao Z, Li Y, Quan X, Zhang Y (2017b) Towards engineering application: potential mechanism for
enhancing anaerobic digestion of complex organic waste with different types of conductive
materials. Water Res 115:266–277
2 Engineering Direct Interspecies Electron Transfer for Enhanced. . . 59
Zhao Z, Zhang Y, Li Y, Dang Y, Zhu T, Quan X (2017c) Potentially shifting from interspecies
hydrogen transfer to direct interspecies electron transfer for syntrophic metabolism to resist
acidic impact with conductive carbon cloth. Chem Eng J 313:10–18
Zhao Z, Li Y, He J, Zhang Y (2018) Establishing direct interspecies electron transfer during
laboratory-scale anaerobic digestion of waste activated sludge via biological ethanol-type
fermentation pretreatment. ACS Sustain Chem Eng 6:13066–13077
Zhao Z, Li Y, Zhang Y, Lovley DR (2020a) Sparking anaerobic digestion: promoting direct
interspecies electron transfer to enhance methane production. iScience 23:101794
Zhao Z, Wang J, Li Y, Zhu T, Yu Q, Wang T, Liang S, Zhang Y (2020b) Why do dieters like
drinking: metagenomic analysis for methane and energy metabolism during anaerobic digestion
with ethanol. Water Res 171:115425
Zheng S, Liu F, Wang B, Zhang Y, Lovley DR (2020) Methanobacterium capable of direct
interspecies electron transfer. Environ Sci Technol 54:15347–15354
Zhou N, Wang T, Chen S, Hu Q, Cheng X, Sun D, Vupputuri S, Qiu B, Liu H, Guo Z (2021)
Conductive polyaniline hydrogel enhanced methane production from anaerobic wastewater
treatment. J Colloid Interface Sci 581:314–322
Zhu J, Zheng H, Ai G, Zhang G, Liu D, Liu X, Dong X (2012) The genome characteristics and
predicted function of methyl-group oxidation pathway in the obligate aceticlastic methanogens,
Methanosaeta spp. PLoS One 7:e36756
Zhu Y, Zhao Z, Zhang Y (2019) Using straw as a bio-ethanol source to promote anaerobic digestion
of waste activated sludge. Bioresour Technol 286:121388
Zhuang L, Tang J, Wang Y, Hu M, Zhou S (2015) Conductive iron oxide minerals accelerate
syntrophic cooperation in methanogenic benzoate degradation. J Hazard Mater 293:37–45
Part II
Pretreatment
Chapter 3
Adsorbents for the Detoxification
of Lignocellulosic Wastes Hydrolysates
to Improve Fermentative Processes
to Bioenergy and Biochemicals Production
Abstract The depletion of fossil fuels and their environmental impact has moti-
vated the research on alternative sources of energy. Lignocellulosic wastes are
potential substrates for bioenergy and biochemicals production due to the carbohy-
drate content in the cell wall, which is composed of cellulose (40–80%), hemicel-
lulose (10–40%), and lignin (5–25%). The sugar content of lignocellulosic biomass
can be converted into fermentable monomeric sugars by physicochemical or enzy-
matic process. The hydrolysates produced during those processes have a high
content of inhibitory substances such as phenolic and furanic compounds, i.e.
syringaldehyde, vanillin, furfural and 5 hydroxymethylfurfural. Thus, the need of a
detoxification pretreatment of hydrolysates before the fermentation process should
be addressed. There are several detoxification processes reported such as overliming,
ion exchange, membranes, use of enzymes or microorganisms and adsorption.
Adsorption processes with nanomaterials have emerged as a promising technique
for the removal of such inhibitors. This chapter shows the potential methods to
detoxify hydrolysates, specifically using the adsorption process with potential adsor-
bents such as nanomaterials. Furthermore, inhibitors should not be just considered as
obstacles for fermentation and hydrolysis, but could also be viewed as valuable
chemicals for other industries which is also highlighted in this chapter.
I. Covarrubias-García
División de Ciencias Ambientales, Instituto Potosino de Investigación Científica y Tecnológica,
San Luis Potosí, Mexico
e-mail: [email protected]
S. Arriaga (*)
División de Ciencias Ambientales, Instituto Potosino de Investigación Científica y Tecnológica,
San Luis Potosí, Mexico
Microbiology Department, School of Natural Sciences, National University of Ireland, Galway,
Galway, Ireland
e-mail: [email protected]
3.1 Introduction
The price rise, environmental issues and depletion of fossil fuels along with the
tremendous increase in the world’s energy demand have pushed countries to search
for alternative renewable energy options such as biofuel production from lignocel-
lulosic biomass (Onaran et al. 2020). Lignocellulosic biomass is one of the most
abundant biopolymers found on earth and includes agricultural residues such as
sugarcane bagasse, rice straw, wheat, soy, oats, corn, stump, sawdust, plant and tree
branches as well as urban solid residues. The secondary cell wall of plants contains
cellulose (40–80%), hemicellulose (10–40%), and lignin (5–25%). Lignocellulosic
biomass contain fermentable sugars that have been extensively evaluated for biofuel
production. The carbohydrate fraction of the plant cell wall is converted into
fermentable monomeric sugars through physical (e.g. milling, ultrasound irradiation
and extrusion), chemical (e.g. acid, alkaline, ammonia explosion, organo-solvents
and ionic liquids), physicochemical (e.g. steam explosion, microwave and CO2
explosion) and enzymatic hydrolysis (e.g. cellulose and xylanase) pretreatment or
a combination of these methods.
During the pretreatment of lignocellulosic biomass, formation of many inhibitory
compounds can take place. These compounds are divided in three main groups based
on their origin: weak acids, furan derivatives and phenolic compounds (Palmqvist
and Hahn-Hägerdal 2000a). Formation of these inhibitors during pretreatment is a
limiting factor in the production of valuable chemicals and bioenergy from ligno-
cellulosic biomass and the selection of an efficient pretreatment method can signif-
icantly reduce the production of inhibitory compounds and may lead to enhance in
production and yield of the desired product. Presence of inhibitory compounds have
adverse effect on the substrate utilization and product yield by fermentative
microbes. The unfavourable environment created by the presence of inhibitors
increases the length of the lag phase, reduces cell density and lower the growth
rates of fermenting microbes (Zabed et al. 2019; Wang et al. 2018; Kumar et al.
2020). Moreover, inhibitors can decrease the activity of several enzymes, break
down the DNA, inhibit protein and RNA synthesis in microorganisms (Jung et al.
2014, 2013; Liu et al. 2004; Palmqvist and Hahn-Hägerdal 2000a, 2000b;
Taherzadeh et al. 2000).
In order to decrease the amount of toxic compounds, several detoxification
methods have emerged such as alkaline detoxification, biological detoxification by
microorganisms or enzymes, and adsorption using polymeric sorbents or activated
carbon (Grzenia et al. 2010; Ludwig et al. 2013; Myoung et al. 2010; Zhang et al.
2011). Adsorption processes are advantageous to detoxify hydrolysates as they do
not modify the chemical composition of the inhibitor which can be desorbed to
3 Adsorbents for the Detoxification of Lignocellulosic Wastes Hydrolysates. . . 65
recover and further utilized as valuable chemical in other industries (Sjulander and
Kikas 2020).
This book chapter is focused on examining the fundamentals that govern the
function of adsorbents for the detoxification of lignocellulosic wastes hydrolysates to
improve fermentative processes to bioenergy and biochemicals production. It also
reviews the commonly used adsorbents and discusses the advantages and disadvan-
tages of using each of these methods. Likewise, it will also present the emerging
methods used for detoxification and the challenges involved in their application.
Physical pretreatment includes biomass particle size reduction using grinding, cut-
ting, ball milling among other processes (Table 3.1). The advantage of this process is
that it does not produce inhibitory substances. Particle size reduction is necessary
prior to most physicochemical pretreatments to improve material handling and
enhances the efficiency of physicochemical pretreatments. However, the excessive
reduction of size makes the processes economically unfeasible for a potential
application, its inability to remove lignin is the main drawback (Taylor et al.
2019). Other physical methods include ultrasonic, extrusion, microwave irradiation,
pulsed electric field, freeze pretreatment and pyrolysis.
66 I. Covarrubias-García and S. Arriaga
Table 3.1 Physical, chemical and biological methods to pretreat lignocellulosic biomass (Bajpai
2016; Bhatia et al. 2020; Bhutto et al. 2017; Zabed et al. 2019)
Type of pretreatment Advantages Disadvantages
Physical • Milling • Increases surface area • Inability to remove
• Ultrasound • Decrystallizes cellulose lignin
• Gamma irradi- • Removes hemicellulose • High energy
ation • No formation of inhibitors requirement
• Extrusion • Improves hydrolysis when used
• Microwave in combination with other
radiation processes
• Hydrothermal
pyrolysis
Chemical • Acid • Solubilize hemicellulose • Inhibitors formation at
hydrolysis • Removes lignin high acid concentration
• Improves sugar yield • Causes corrosion of
equipment
• The acid recycling
process is expensive
• Requires neutralization
• Alkaline • Solubilize lignin and hemicel- • Reaction time takes
hydrolysis lulose days-weeks at room
• Increase crystallinity of cellu- temperature
lose • Harsh chemical
• Improves glucose yield from conditions
biomass
• Organosolv • Efficient lignin removal • Recycling of solvent
process • Pure lignin obtention • Risk in operating at
• High sugar yield from biomass high pressure
• Recovery of lignocellulosic
components
• Ionic liquids • Effective in lignin removal • Extremely viscous
(IL) • Decrystallize cellulose solution is formed which
• As they are non-volatile, bio- limit industrial applica-
mass is easily treated at high tion
temperature without solvent • Complex synthesis
losses by evaporation process
• ILs can be recovered and • High costs of ILs
recycled easily
• Ozonolysis • Selective lignin degradation • High cost of generating
• Produce biomass with great ozone
surface areas
• Minimal losses of hemicellu-
lose and cellulose
• Peroxide • Depolymerize lignin • Requires to maintain
• Mild conditions of temperature the pH constant to avoid
and pressure hemicellulose
• Reduce inhibitors formation elimination
Combined • Steam • Solubilize hemicellulose • Process cost is high
physical and explosion • High removal of lignin • High pressure required
chemical • Increase porosity and surface • High rate of degrada-
methods area by disruption of biomass tion of sugars
(continued)
3 Adsorbents for the Detoxification of Lignocellulosic Wastes Hydrolysates. . . 67
is changed. This pretreatment is highly selective for the biomass constituents and
requires harsh operational conditions (Peral 2016). These methods include acid
hydrolysis, alkaline hydrolysis, organosolv, peroxide and ozonolysis (Table 3.1).
The most known acids used are H2SO4, HNO4, HCl, H3PO4, HNO3 and organic
acids.
Acid hydrolysis is considered one of the most common lignocellulosic biomass
pretreatments and can be performed by using dilute or concentrated acids (4% to
40%). They are flexible in feedstock choice, high monomeric sugar yield, and mild
temperature operation conditions. The formation of toxic compounds such as
5 hydroxymethylfurfural (5-HMF), furfural, levulinic and formic acid occurs under
acid hydrolysis. A neutralization step is required which adds cost to the overall
process.
During alkaline treatment, little degradation of sugars to furfural, HMF and
organic acids occurs (Peral 2016). Organic solvent pretreatment involves the use
of an organic or aqueous organic solvent mixed with inorganic acid catalysts (HCl or
H2SO4). Oxalic, acetylsalicylic and salicylic acid can also be used as catalysts.
Solvents commonly used are methanol, ethanol, acetone and ethylene glycol (Peral
2016). Organosolv pretreatment hydrolyses lignin bonds as well as lignin-
carbohydrate bonds. Lignin is extensively removed and hemicellulose is almost
completely solubilized, while cellulose remains in solid form (Peral 2016). This
pretreatment is very selective to lignin which can be recovered and utilized for a
range of applications such as precursors for various chemicals or fuels in a
biorefinery. The challenge with this pretreatment is the recovery of the organic
solvents. The high cost involved in the organosolv pretreatment method due to
involvement of expensive chemicals and catalyst is the only drawback of this
method which make its application less desirable than the other commonly used
technologies (Peral 2016). Moreover, in many cases, high temperature (up to
200 C) is also applied in organosolv pretreatment which can increase the cost of
operation.
liquids are effective in decrystallizing the cellulose, fractionating lignin and poly-
saccharide contents into separate streams, making the following hydrolysis step
much easier (Aslanzadeh et al. 2014).
Fig. 3.1 Inhibitory compounds profile derived from lignocellulosic materials during pretreatment
et al. 2017; Yu et al. 2011). Table 3.2 shows the common inhibitors that can be
formed from pretreatment of lignocellulosic biomass and their principal effects on
fermentative microorganisms.
Table 3.2 and Fig. 3.1 show that inhibitors come from the degradation of each
component of the biomass. Cellulose can degrade to hexoses to different extent.
Such hexoses could further be dehydrated to 5-HMF, and 5-HMF can be further
dehydrated to form levulinic acid and formic acid (Kumar et al. 2018). On the other
hand, hemicellulose after pretreatment can degrade into sugar acids, aliphatic acids,
and furan aldehydes, where furfural is an abundant and potent inhibitor (Osorio-
González et al. 2019; Kumar et al. 2020). Formic acid, acrylic acid and levulinic acid
are other carboxylic acids found in hemicellulose hydrolysates (Zabed et al. 2019;
Kumar et al. 2020).
Regarding the degradation products from lignin, 4-hydroxybenzoic acid,
4-hydroxybenzaldehyde, vanillin, dihydroconiferyl alcohol, coniferyl aldehyde,
syringaldehyde and syringic acid are the most common phenols formed from lignin
degradation (Hodge et al. 2009; Kumar et al. 2020). There are different effects
associated with each inhibitor. For instance, weak acids in undissociated form can
permeate through and inside the cells to release the anion and proton disrupting the
intracellular pH (Wang et al. 2018). On the other hand, phenolics are more toxic than
aliphatic and furans with the same functional groups and have shown antibiofouling
3 Adsorbents for the Detoxification of Lignocellulosic Wastes Hydrolysates. . . 71
effects on Gram negative bacteria (Pattrick et al. 2019). Likewise, furans present
negative effects on glycolytic and fermentative enzymes and can inhibit the synthe-
sis of ATP, which lead to the cessation of DNA replication (Basak et al. 2020).
Furfural and 5-HMF bind to nitrogenous bases of DNA and induce strand break,
damaging the structure (Liu et al. 2019). Both phenolic compounds and furans
generate Reactive Oxygen Species (ROS), which inactivate enzymes, damaging
DNA and cellular proteins (Basak et al. 2020; Lin et al. 2015).
In general, the effect of each inhibitor and the degree of inhibition will depend on
the fermenting microorganisms and its tolerance to the inhibitor. Additionally, the
toxicity of each inhibitor depends on its concentration, chemical structure and
reactivity, molecular weight and chemical polarity, which determines their inherent
capacity to penetrate the cell membrane and cause cellular damage (Jung et al. 2013;
Taherzadeh et al. 2000).
In general if there is more than one inhibitory compounds are present in the
lignocellulosic hydrolysates that can have a more severe recalcitrant effect on the
microorganisms and enzymatic activity (Almeida et al. 2007; Jayakody et al. 2017;
72 I. Covarrubias-García and S. Arriaga
Wikandari et al. 2019; Sjulander and Kikas 2020). For instance, the effects of
inhibitor combinations (i.e. acetate, furfural, 5-HMF, vanillin,
p-hydroxybenzaldehyde and syringaldehyde) on R. toruloides-Y4 at different con-
centrations have been analysed. The combinations of the different inhibitors gave a
much more complex inhibition effect on the fermentation process (Hu et al. 2009;
Zhao et al. 2012).
Currently, different strategies have been employed to tackle problems with
inhibition after pretreatment of lignocellulose. Some of the different approaches
used are: (i) feedstock selection, that is using less recalcitrant feedstocks in order to
generate less inhibitors during pretreatment (Chiaramonti et al. 2012; Larsen et al.
2012); (ii) selection of microorganisms, i.e. the selection of microorganisms that
produce less amount of inhibitors, should be done mainly based on specific produc-
tivity and product yields (Wimalasena et al. 2014); (iii) adaptive evolution, where
microorganism adapt to specific inhibitors (Almario et al. 2013; Koppram et al.
2012); (iv) genetic or metabolic engineering, based on creating tolerance to fermen-
tation inhibitors (Sanda et al. 2011; Wang et al. 2013); (v) culturing schemes such as
the use of a large inoculum size (Hoyer et al. 2010; Pienkos and Zhang 2009);
(vi) conditioning which consists of addition of chemicals that can improve the
fermentability of the substrates (e.g. NH4OH and NaOH), these agents can be
added during fermentation (Alriksson et al. 2011, 2006); and (vii) detoxification of
hydrolysate (e.g. distillation, solvent extraction and adsorption) (Mussatto and
Roberto 2001; Myoung et al. 2010; Palmqvist and Hahn-Hägerdal 2000a).
Section 3.4 described that removal of inhibitors from hydrolysate is an essential step
before performing microbial fermentation. Table 3.3 overviews different adsorption
materials that have been tested for their ability to remove inhibitors from lignocel-
lulosic hydrolysates. The types of adsorbents include activated carbons, minerals,
resins, industrial and agricultural wastes. The adsorption efficiency of the inhibitors
depends on the adsorbent surface area, pore size and chemical properties, and is also
influenced by environmental conditions (i.e. temperature, pH, and contact time
between inhibitor and adsorbent) (Deng et al. 2018; Sjulander and Kikas 2020).
Activated carbon (AC) either stand-alone or in combination with other chemicals
(i.e. alkali and ion exchange resins) is the most common adsorbent used for detox-
ification. Activated carbons are usually produced from low-cost materials which is
one of its main advantage. Apart from that, reliability and consistency of the resource
supply, ease of activation, adsorption capacity, and selectivity are notable qualities
of AC. Literature has reported several studies where different inhibitors have been
removed with activated carbons, such as furfural, 5-HMF, acetic acid and phenolic
compounds (Chen et al. 2019; Myoung et al. 2010; Santana et al. 2018; Zhang et al.
2011). It has been suggested that removal of inhibitors is related to the strong
hydrophobic nature of the activated carbons (Lee and Park 2016). The main draw-
backs in the use of AC are its regenerability and cost, where there is 10% loss during
each regeneration cycle. In addition, it can retain up 30% of the fermentable sugars
which is not good from the economic point of view (Carvalho et al. 2006; Ranjan
et al. 2009). Therefore, the exploration and study of other adsorbents is of interest.
Alternatively, mineral adsorbents including siliceous materials, clay and natural
zeolites have been used (Wang and Peng 2010). Ranjan et al. (2009) demonstrated
the potential recovery of HMF, furfural and vanillin using hydrophobic zeolites.
Removal of model compounds using mineral adsorbents has been also reported
(Ahmaruzzaman 2008). On the other hand, different types of resins have been
reported to remove inhibitors such as fumaric acid, acetic acid, formic acid,
5-HMF and soluble lignin (Chen et al. 2019; Choi et al. 2017; Zheng et al. 2018).
The main advantages in the use of polymeric resins is that they are durable,
chemically inert and stable, and possess a high adsorption capacity, efficiency,
selectivity and ease of regeneration, with relatively low cost and limited toxicity
(Soto et al. 2011).
In addition, the implementation of nanoscale materials has emerged as an attrac-
tive alternative in several applications. Their size and their relatively large surface
area to volume ratio when compared to larger particles can result in a high adsorption
capacity (Alsaba et al. 2020). Few studies have reported the use of nanomaterials for
detoxification of hydrolysates. Covarrubias-García et al. (2021) used reduced
graphene oxide adorned with magnetite nanoparticles for the removal of fermenta-
tion inhibitors such as furfural, 5-HMF, levulinic acid, vanillic acid and vanillin from
3 Adsorbents for the Detoxification of Lignocellulosic Wastes Hydrolysates. . . 75
Table 3.3 Summary of adsorbents used for detoxification of hydrolysates. “x” means the studies
where fermentation step is not performed to produce any product
Removal of
inhibitors
Type Adsorbent (%) Products Reference
Activated Activated carbon HMF (96) x Myoung et al.
carbons Furfural (2010)
(93)
Formic acid
(42)
Acetic acid
(14)
Activated carbon Furfural Ethanol Zhang et al.
(97) (2011)
Activated charcoal Phenolic Xylitol Santana et al.
compounds (2018)
(78)
Furfural
(99)
5-HMF (99)
Activated carbon HMF Acetone, Wang et al.
(50–60) butanol and (2019)
Furfural ethanol
(50–60)
Phenolic
compounds
(50)
Membranes Adsorptive membrane Acetic acid x Wickramasinghe
(Sartobind Q) (99) and Grzenia
(2008)
Hollow-structured porous 5-HMF (94) x Zhang et al.
aromatic polymer (2019)
Resins Amberchrom-CG71C resin Fumaric x Choi et al. (2017)
acid (99)
Acetic acid
(99)
Microporous polymeric Levulinic x Lin et al. (2017)
SY-01 resin acid (99)
Microporous resin (SY-01) Formic acid x Zheng et al.
(80) (2018)
Levulinic
acid (16)
5-HMF (63)
Resin CS-6 Soluble lig- Lipid Chen et al.
nin (70) (2019)
Nanostructures Fe0 nanoparticles on acti- Furfural x Sajab et al.
vated carbon (99) (2019)
Acid solu-
ble lignin
(81)
(continued)
76 I. Covarrubias-García and S. Arriaga
wood hydrolysates with a minimal sugar loss of 5.9, 6.2 and 7.6%, for 10, 20 and
50 mg of the adsorbent, respectively. There are other studies where nanomaterials
are immobilized on the AC surface (Sajab et al. 2019) and nanomaterials as a support
for enzyme immobilization (Yin et al. 2020) have been reported for removal of
toxicants from hydrolysate.
Other types of adsorbents such as sugarcane bagasse derived fly ash has been
successfully used to adsorb aromatic compounds, such as tannic acid and vanillin
from a lignocellulosic hydrolysate (Freitas and Farinas 2017).
The main challenge related to the adsorbent based technology for detoxification of
lignocellulosic hydrolysates is related to the desorption and reusability of the
adsorbents. Desorbing eluents may not be effective for all the adsorbents and need
3 Adsorbents for the Detoxification of Lignocellulosic Wastes Hydrolysates. . . 77
point extraction (CPE) has also been explored. Surfactants, which are soluble in
water at low temperatures but insoluble at high temperatures, are used in this
approach, allowing for a two-phase extraction. The main advantage of the method
is that the surfactants are non-toxic to microorganisms and no sugar loss has been
observed (Dhamole et al. 2013). Other studies have reported the extraction of
inhibitors with common solvents like deionized water and ethanol. Ranjan et al.
(2009) demonstrated the potential recovery of HMF, furfural and vanillin using
hydrophobic zeolites. They desorbed with deionized water and indicated that desorp-
tion isotherms almost overlapped with the adsorption isotherms. Lin et al. (2017)
found that the use of an ethanol solution as an eluent was easy and effective to
recover levulinic acid from a SY-01 resin with a 99.4% efficiency and the SY-01
resin was successfully regenerated.
References
Liu ZL, Slininger PJ, Dien BS, Berhow MA, Kurtzman CP, Gorsich SW (2004) Adaptive response
of yeasts to furfural and 5-hydroxymethylfurfural and new chemical evidence for HMF con-
version to 2, 5-bis-hydroxymethylfuran. J Ind Microbiol Biotechnol 31:345–352. https://doi.
org/10.1007/s10295-004-0148-3
Liu H, Zhang J, Yuan J, Jiang X, Jiang L, Zhao G, Huang D, Liu B (2019) Omics-based analyses
revealed metabolic responses of Clostridium acetobutylicum to lignocellulose-derived inhibitors
furfural, formic acid and phenol stress for butanol fermentation. Biotechnol Biofuels 12:101.
https://doi.org/10.1186/s13068-019-1440-9
Llano T, Quijorna N, Coz A (2017) Detoxification of a lignocellulosic waste from a pulp mill to
enhance its fermentation prospects. Energies 10(3):348. https://doi.org/10.3390/en10030348
Ludwig D, Amann M, Hirth T, Rupp S, Zibek S (2013) Bioresource technology development and
optimization of single and combined detoxification processes to improve the fermentability of
lignocellulose hydrolyzates. Bioresour Technol 133:455–461. https://doi.org/10.1016/j.
biortech.2013.01.053
Lynd L, Weimer P, van Zyl W, Pretorius I (2002) Microbial cellulose utilization: fundamentals and
biotechnology. Microbiol Mol Biol Rev 66:506–577
Mussatto SI, Roberto IC (2001) Hydrolysate detoxification with activated charcoal for xylitol
production by Candida guilliermondii. Biotechnol Lett 23:1681–1684. https://doi.org/10.
1023/A:1012492028646
Mussatto SI, Dragone G, Fernandes M, Milagres AMF, Roberto IC (2008) The effect of agitation
speed, enzyme loading and substrate concentration on enzymatic hydrolysis of cellulose from
brewer’s spent grain. Cellulose 15:711. https://doi.org/10.1007/s10570-008-9215-7
Myoung J, Venditti RA, Jameel H, Kenealy WR (2010) Detoxification of woody hydrolyzates with
activated carbon for bioconversion to ethanol by the thermophilic anaerobic bacterium
Thermoanaerobacterium saccharolyticum. Biomass Bioenergy 35:626–636. https://doi.org/
10.1016/j.biombioe.2010.10.021
Nilvebrant N-O, Persson P, Reimann A, de Sousa F, Gorton L, Jönsson LJ (2003) Limits for
alkaline detoxification of dilute-acid lignocellulose hydrolysates. Appl Biochem Biotechnol
107:615–628. https://doi.org/10.1385/ABAB:107:1-3:615
Onaran G, Gürel L, Argun H (2020) Detoxification of waste hand paper towel hydrolysate by
activated carbon adsorption. Int J Environ Sci Technol 17:799–808. https://doi.org/10.1007/
s13762-019-02499-w
Osorio-González CS, Hegde K, Brar SK, Kermanshahipour A, Avalos-Ramírez A (2019) Chal-
lenges in lipid production from lignocellulosic biomass using Rhodosporidium sp.; a look at the
role of lignocellulosic inhibitors. Biofuels Bioprod Biorefin 13:740–759. https://doi.org/10.
1002/bbb.1954
Palmqvist E, Hahn-Hägerdal B (2000a) Fermentation of lignocellulosic hydrolysates. I: inhibition
and detoxification. Bioresour Technol 74:17–24. https://doi.org/10.1016/S0960-8524(99)
00160-1
Palmqvist E, Hahn-Hägerdal B (2000b) Fermentation of lignocellulosic hydrolysates. II: inhibitors
and mechanisms of inhibition. Bioresour Technol 74:25–33. https://doi.org/10.1016/S0960-
8524(99)00161-3
Pattrick CA, Webb JP, Green J, Chaudhuri RR, Collins MO, Kelly DJ (2019) Proteomic profiling,
transcription factor modeling, and genomics of evolved tolerant strains elucidate mechanisms of
vanillin toxicity in Escherichia coli. mSystems 4:e00163–e00119. https://doi.org/10.1128/
mSystems.00163-19
Peral C (2016) Biomass pretreatment strategies (technologies, environmental performance, eco-
nomic considerations, industrial implementation). In: Biotransformation of agricultural waste
and by-products. Elsevier, San Diego, pp 125–160. https://doi.org/10.1016/B978-0-12-
803622-8.00005-7
Pienkos PT, Zhang M (2009) Role of pretreatment and conditioning processes on toxicity of
lignocellulosic biomass hydrolysates. Cellulose 16:743–762. https://doi.org/10.1007/s10570-
009-9309-x
82 I. Covarrubias-García and S. Arriaga
Poontawee R, Yongmanitchai W, Limtong S (2017) Efficient oleaginous yeasts for lipid production
from lignocellulosic sugars and effects of lignocellulose degradation compounds on growth and
lipid production. Process Biochem 53:44–60. https://doi.org/10.1016/j.procbio.2016.11.013
Qi B, Luo J, Chen X, Hang X, Wan Y (2011) Separation of furfural from monosaccharides by
nanofiltration. Bioresour Technol 102:7111–7118. https://doi.org/10.1016/j.biortech.2011.
04.041
Ranjan R, Thust S, Gounaris CE, Woo M, Floudas CA, von Keitz M, Valentas KJ, Wei J, Tsapatsis
M (2009) Adsorption of fermentation inhibitors from lignocellulosic biomass hydrolyzates for
improved ethanol yield and value-added product recovery. Microporous Mesoporous Mater
122:143–148. https://doi.org/10.1016/j.micromeso.2009.02.025
Sajab MS, Santanaraj J, Mohammad AW, Kaco H, Harun S (2019) Detoxification of lignocellulosic
hydrolysates by in situ formation of Fe (0) nanoparticles on activated carbon. Bio Resources 14:
8614–8626. https://doi.org/10.15376/biores.14.4.8614-8626
Sanda T, Hasunuma T, Matsuda F, Kondo A (2011) Repeated-batch fermentation of lignocellulosic
hydrolysate to ethanol using a hybrid Saccharomyces cerevisiae strain metabolically engineered
for tolerance to acetic and formic acids. Bioresour Technol 102:7917–7924. https://doi.org/10.
1016/j.biortech.2011.06.028
Santana NB, Dias JCT, Rezende RP, Franco M, Oliveira LKS, Souza LO (2018) Production of
xylitol and bio-detoxification of cocoa pod husk hemicellulose hydrolysate by Candida boidinii
XM02G. PLoS One 13:e0195206
Singh B, Verma AP, Mandal PK, Datta S (2017) A biotechnological approach for degradation of
inhibitory compounds present in lignocellulosic biomass hydrolysate liquor using Bordetella
sp. BTIITR Chem Eng J 328:519–526. https://doi.org/10.1016/j.cej.2017.07.059
Sjulander N, Kikas T (2020) Origin, impact and control of lignocellulosic inhibitors in bioethanol
production—a review. Energies 13(18):4751. https://doi.org/10.3390/en13184751
Soto ML, Moure A, Domínguez H, Parajó JC (2011) Recovery, concentration and purification of
phenolic compounds by adsorption: a review. J Food Eng 105:1–27. https://doi.org/10.1016/j.
jfoodeng.2011.02.010
Suman SK, Khatri M, Dhawaria M, Kurmi A, Pandey D, Ghosh S, Jain SL (2018) Potential of
Trametes maxima IIPLC-32 derived laccase for the detoxification of phenolic inhibitors in
lignocellulosic biomass prehydrolysate. Int Biodeterior Biodegradation 133:1–8. https://doi.org/
10.1016/j.ibiod.2018.05.009
Taherzadeh MJ, Gustafsson L, Niklasson C, Lidén G (2000) Physiological effects of
5-hydroxymethylfurfural on Saccharomyces cerevisiae. Appl Microbiol Biotechnol 53:701–
708. https://doi.org/10.1007/s002530000328
Taylor MJ, Alabdrabalameer HA, Skoulou V (2019) Choosing physical, physicochemical and
chemical methods of pre-treating lignocellulosic wastes to repurpose into solid fuels. Sustain-
ability 11(13):3604. https://doi.org/10.3390/su11133604
Tomek KJ, Saldarriaga CRC, Velasquez FPC, Liu T, Hodge DB, Whitehead TA (2015) Removal
and upgrading of lignocellulosic fermentation inhibitors by in situ biocatalysis and liquid-liquid
extraction. Biotechnol Bioeng 112:627–632. https://doi.org/10.1002/bit.25473
Travália BM, Santos NTDG, Vieira MGA, Forte MBS (2019) Adsorption of fermentation inhibitors
by layered double hydroxides in synthetic hemicellulose hydrolysate: a batch multicomponent
analysis. Ind Eng Chem Res 58:18822–18828. https://doi.org/10.1021/acs.iecr.9b03184
Wang S, Peng Y (2010) Natural zeolites as effective adsorbents in water and wastewater treatment.
Chem Eng J 156:11–24. https://doi.org/10.1016/j.cej.2009.10.029
Wang X, Yomano LP, Lee JY, York SW, Zheng H, Mullinnix MT, Shanmugam KT, Ingram LO
(2013) Engineering furfural tolerance in Escherichia coli improves the fermentation of ligno-
cellulosic sugars into renewable chemicals. Proc Natl Acad Sci 110:4021–4026. https://doi.org/
10.1073/pnas.1217958110
Wang S, Sun X, Yuan Q (2018) Strategies for enhancing microbial tolerance to inhibitors for
biofuel production: a review. Bioresour Technol 258:302–309. https://doi.org/10.1016/j.
biortech.2018.03.064
3 Adsorbents for the Detoxification of Lignocellulosic Wastes Hydrolysates. . . 83
Wang P, Chen YM, Wang Y, Lee YY, Zong W, Taylor S, McDonald T, Wang Y (2019) Towards
comprehensive lignocellulosic biomass utilization for bioenergy production: efficient
biobutanol production from acetic acid pretreated switchgrass with clostridium
saccharoperbutylacetonicum N1-4. Appl Energy 236:551–559. https://doi.org/10.1016/j.
apenergy.2018.12.011
Wickramasinghe SR, Grzenia DL (2008) Adsorptive membranes and resins for acetic acid removal
from biomass hydrolysates. Desalination 234:144–151. https://doi.org/10.1016/j.desal.2007.
09.080
Wikandari R, Sanjaya AP, Millati R, Karimi K, Taherzadeh MJ (2019) Chapter 20: Fermentation
inhibitors in ethanol and biogas processes and strategies to counteract their effects. In:
Pandey A, Larroche C, Dussap C-G, Gnansounou E, Khanal SK, Ricke SBT (eds) Biomass,
biofuels, biochemicals. Academic, Amsterdam, pp 461–499. https://doi.org/10.1016/B978-0-
12-816856-1.00020-8
Wimalasena TT, Greetham D, Marvin ME, Liti G, Chandelia Y, Hart A, Louis EJ, Phister TG,
Tucker GA, Smart KA (2014) Phenotypic characterisation of Saccharomyces spp. yeast for
tolerance to stresses encountered during fermentation of lignocellulosic residues to produce
bioethanol. Microb Cell Factories 13:47. https://doi.org/10.1186/1475-2859-13-47
Yin L, Chen J, Wu W, Du Z, Guan Y (2020) Immobilization of laccase on magnetic nanoparticles
and application in the detoxification of rice straw hydrolysate for the lipid production of
Rhodotorula glutinis. Appl Biochem Biotechnol 193(4):998–1010. https://doi.org/10.1007/
s12010-020-03465-w
Yu X, Zheng Y, Dorgan KM, Chen S (2011) Oil production by oleaginous yeasts using the
hydrolysate from pretreatment of wheat straw with dilute sulfuric acid. Bioresour Technol
102:6134–6140. https://doi.org/10.1016/j.biortech.2011.02.081
Zabed HM, Akter S, Yun J, Zhang G, Awad FN, Qi X, Sahu JN (2019) Recent advances in
biological pretreatment of microalgae and lignocellulosic biomass for biofuel production.
Renew Sust Energ Rev 105:105–128. https://doi.org/10.1016/j.rser.2019.01.048
Zevallos Torres LA, Lorenci Woiciechowski A, de Andrade Tanobe VO, Karp SG, Guimarães
Lorenci LC, Faulds C, Soccol CR (2020) Lignin as a potential source of high-added value
compounds: a review. J Clean Prod 263:121499. https://doi.org/10.1016/j.jclepro.2020.121499
Zhang Y-HP (2011) What is vital (and not vital) to advance economically-competitive biofuels
production. Process Biochem 46:2091–2110. https://doi.org/10.1016/j.procbio.2011.08.005
Zhang K, Agrawal M, Harper J, Chen R, Koros WJ (2011) Removal of the fermentation inhibitor,
furfural, using activated carbon in cellulosic-ethanol production. Ind Eng Chem Res
50(24):14055–14060. https://doi.org/10.1021/ie2013983
Zhang Y-B, Luo Q-X, Lu M-H, Luo D, Liu Z-W, Liu Z-T (2019) Controllable and scalable
synthesis of hollow-structured porous aromatic polymer for selective adsorption and separation
of HMF from reaction mixture of fructose dehydration. Chem Eng J 358:467–479. https://doi.
org/10.1016/j.cej.2018.10.029
Zhao X, Peng F, Du W, Liu C, Liu D (2012) Effects of some inhibitors on the growth and lipid
accumulation of oleaginous yeast Rhodosporidium toruloides and preparation of biodiesel by
enzymatic transesterification of the lipid. Bioprocess Biosyst Eng 35:993–1004. https://doi.org/
10.1007/s00449-012-0684-6
Zheng J, Pan B, Xiao J, He X, Chen Z, Huang Q, Lin X (2018) Experimental and mathematical
simulation of noncompetitive and competitive adsorption dynamic of formic acid–Levulinic
acid–5-Hydroxymethylfurfural from single, binary, and ternary systems in a fixed-bed column
of SY-01 resin. Ind Eng Chem Res 57:8518–8528. https://doi.org/10.1021/acs.iecr.8b01283
Chapter 4
Pretreatment of Lignocellulosic Materials
to Enhance their Methane Potential
Abstract Lignocellulosic materials (LMs) are the most abundant residues on the
planet and have a huge potential for methane production. Several strategies have
been tested to enhance the methane potential of LMs, with a particular emphasis on
environmentally friendly and economically convenient pretreatments. This chapter
revisits the potential of two chemical, i.e. organosolv and N-methylmorpholine
N-oxide (NMMO)-driven, and one physical, i.e. ultrasounds, pretreatments.
Organosolv pretreatment enables to obtain a pure lignin fraction from LMs, leaving
most of the fermentable sugars in the solid matrix. The result is a lignin-poor material
with an increased porosity and a higher bioavailability of the sugar fraction. Another
advantage is the cost-effectiveness and the easy recovery of the chemicals involved.
NMMO pretreatment focuses on the cellulosic component of the biomass, aiming to
reduce its crystallinity and to increase the porosity of the substrate. The main
advantage of NMMO lies in its high recovery percentage, which reaches up to
99%. Ultrasound pretreatment involves ultrasonic waves that allow fractionating
LMs, breaking the linkages between lignin, cellulose and hemicellulose, generally
leaving cellulose and most of the hemicellulose in the solid fraction and dissolving
the lignin in the liquor. Ultrasound pretreatment does not require chemicals and can
be easily combined with other pretreatment methods to enhance its effectiveness.
4.1 Introduction
The interest in renewable energy started when the world experienced an energy crisis
in the ‘70s and realised that fossil fuels are not an unlimitedly available resource.
More recently, the interest in renewable energy sources is related to global environ-
mental quality (Sen and Ganguly 2017) and geographic independence of the fossil
fuels reservoirs (Martins et al. 2019). The first emerging problem regarded the
emission of toxic compounds and oxides of nitrogen and sulfur upon combustion
of fossil fuels, which contributes to acid rain formation (Singh and Agrawal 2008).
At the moment, the main concern relates to global warming caused by the increase in
carbon dioxide concentration and other upper atmospheric pollutants deriving from
anthropogenic activities (Kweku et al. 2018; Qazi et al. 2019).
The preservation of the environment and the necessity of renewable energy to
replace fossil fuels has led to the development of a strong anaerobic digestion
(AD) infrastructure worldwide (Vasco-Correa et al. 2018). AD allows for recovering
energy (i.e. biogas, a CH4/CO2 mixture) from wastes while stabilizing the residual
solid/liquid organic fraction (Bharathiraja et al. 2018). AD involves different com-
munities of microorganisms in four stages: hydrolysis, acidogenesis, acetogenesis,
and methanogenesis (Chen et al. 2008; Zhang et al. 2014). During the first stage,
hydrolytic bacteria secrete enzymes able to hydrolise the organic matter and decom-
pose the complex organic polymers (i.e. lipids, carbohydrates, and proteins) into
soluble monomers (i.e. monomeric sugars, long-chain fatty acids, and amino acids)
and hydrogen. Acidogenesis is the second stage, wherein the soluble monomers are
fermented into volatile fatty acids and other products, which, during the subsequent
acetogenesis stage, are converted to acetate, carbon dioxide and hydrogen by
acetogenic bacteria. In the final phase, methanogenesis, acetic acid, and hydrogen
are converted into methane by methanogenic archaea (Bianco et al. 2021a; Luo et al.
2019).
Many waste materials from several activities are lignocellulosic materials (LMs).
The disposal of these LMs is often difficult due to the enormous volumes produced,
especially during the harvesting season (Barbu et al. 2020; Oh et al. 2018). AD is a
viable and green alternative to landfill disposal or combustion for these LMs
(Alonso-Fariñas et al. 2020). The use of LMs for AD is, however, still limited by
their resistance to biological and chemical degradation. The main issue of AD of
LMs is the complex and resistant structure, mainly consisting of cellulose, hemicel-
lulose, and lignin (Bhatia et al. 2020). Various strategies have been explored to
increase the biodegradability of LMs (Bianco et al. 2021b; Kohli et al. 2020;
Mancini et al. 2018c; Papirio 2020; Yao et al. 2018). Pretreatments of LMs aim to
increase the efficacy of lignocellulose hydrolysis by improving the accessibility of
microorganisms to the sugar fraction (cellulose and hemicellulose) of LMs. This can
be achieved by removing lignin and/or hemicellulose or by decreasing the degree of
polymerization and crystallinity of the cellulosic component of the biomass (Kumar
and Sharma 2017).
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 87
LMs may be converted to various forms of energy, including heat (via burning),
steam, electricity, hydrogen, ethanol, methanol, or methane. The choice depends on
multiple factors, such as conversion efficiencies, transport of energy, need for heat or
steam, economies of scale, and environmental impact. Under most circumstances,
methane is an ideal fuel. Firstly, compared to other fossil fuels, methane produces
few atmospheric pollutants and generates less carbon dioxide per unit of energy.
Therefore, its use for appliances, vehicles, industrial applications, and power gener-
ation is increasing. Furthermore, methane can be used at various stages of purity, and
its transport cost and energy conversion efficiency are comparable to those of
electricity (Chynoweth et al. 2000; Gür 2016). Besides, an extensive pipeline
distribution system is already in place worldwide for methane use. In contrast,
other fuels such as methanol and hydrogen are not well developed commercially
for production, distribution and usage (Chynoweth et al. 2000).
This chapter aims to overview the characteristics and potential of LMs for AD,
with a specific focus on three emerging pretreatment technologies to enhance the
conversion of LMs into methane: organosolv, N-methylmorpholine N-oxide
(NMMO), and ultrasound pretreatment. The principles of each pretreatment will
be thoroughly discussed, pointing out advantages and drawbacks. Particular atten-
tion will be given to the effect of the three pretreatments on the physical character-
istics and chemical composition of LMs, focusing on the subsequent valorisation of
the pretreated solid residues through the AD process.
Biomass is one of the most abundant resources on the planet, with a global
production of 2 1011 tons per year (Reddy and Yang 2005). Biomass can be
converted into fuels and provide renewable materials at the same time, offering a
viable alternative to fossil fuels. Biomass is produced via photosynthesis, by fixing
atmospheric carbon dioxide and converting solar energy to chemical energy to build
up the carbon backbone of plant cells (Zhang 2008). LMs represent more than 60%
of the global biomass and serve as a cheap and abundant feedstock (Bilal et al. 2017).
Agricultural, municipal, and industrial activities generate LMs as waste, generally at
low cost (Table 4.1). Also, the use of LMs for biofuel production does not create
conflicts between land use for food and energy production (Kucharska et al. 2018).
Depending on the origin, LMs are classified as forest residues, municipal solid
waste, waste paper, or crop residue resources (Balat 2011).
The structural and chemical composition of LMs is extremely variable (Table 4.1)
due to various genetic and environmental factors. The chemical composition of LMs
includes mainly cellulose, hemicellulose, and lignin, arranged in a three-dimensional
and complex structure (Fig. 4.1) (Zhang et al. 2019). Depending on the specific
substrate, a significant part of LMs may consist of non-structural compounds. The
most common extractives present in LMs are free sugars (sucrose, glucose, and
88 A. Oliva et al.
Table 4.1 Chemical composition (in terms of cellulose, hemicellulose, and lignin content) of the
most employed lignocellulosic materials for methane production
Biomass Cellulosea Hemicelluloseb Ligninc
origin Substrate (%) (%) (%) Reference
Agricultural Maize straw 38.3 29.8 3.8 Khatri et al. (2015)
residues Wheat straw 31.0 18.4 18.3 Mancini et al. (2018a)
Rice straw 28.6 19.5 17.3 Mancini et al.
(2018b)
Barley straw 39.1 25.7 15.2 Duque et al. (2013)
Sweet sor- 37.7 28.1 21.5 Dong et al. (2019)
ghum straw
Oat straw 35.0 28.2 4.1 Gomez-Tovar et al.
(2012)
Rye straw 42.1 23.8 19.5 Ingram et al. (2011)
Triticale straw 33.0 23.0 29.0 Teghammar et al.
(2012)
Sugarcane 47.6 22.6 27.6 Hashemi et al.
bagasse (2019a)
Sunflower 34.1 26.2 26.8 Hesami et al. (2015)
stalks
Nuts Peanut shell 23.6 12.2 40.0 Shen et al. (2018)
residues Almond shell 23.4 21.9 30.6 Oliva et al. (2021)
Walnut shell 25.6 23.0 46.7 Şenol (2021)
Pistachio shell 20.1 23.2 24.3 Shen et al. (2018)
Chestnut shell 26.8 24.5 36.8 Bianco et al. (2021b)
Hazelnut shell 18.0 17.2 39.1 Shen et al. (2018)
Hazelnut skin 10.2 3.6 39.7 Oliva et al. (2021)
Industrial Spent coffee 8.8 33.6 20.3 Oliva et al. (2021)
wastes grounds
Brewery spent 19.2 26.9 30.5 Ravindran et al.
grain (2018)
Cocoa bean 13.5 7.0 29.9 Mancini et al.
shell (2018b)
Rubber wood 43.6 8.3 31.0 Tongbuekeaw et al.
waste (2020)
Oil palm 36.1 22.4 26.4 Tang et al. (2018)
empty fruit
bunch
Olive pomace 12.3 8.9 34.0 Elalami et al. (2020)
Grape pomace 15.8 8.6 35.4 Bordiga et al. (2019)
Forest Switch grass 42.0 19.0 24.0 Larnaudie et al.
residues (2019)
Spruce wood 42.0 20.0 27.0 Teghammar et al.
(2012)
Poplar wood 49.0 23.0 27.0 Rego et al. (2019)
Birch wood 40.1 26.8 24.5 Goshadrou et al.
(2013)
(continued)
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 89
4.2.1 Cellulose
Fig. 4.1 Schematic representation of the structure of lignocellulosic materials containing cellulose, hemicellulose, and lignin
A. Oliva et al.
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 91
4.2.2 Hemicellulose
Hemicellulose does not have a fixed structure. Its backbone can be either a homo-
polymer or a hetero-polymer with short branches linked by β-1,4-glucan bonds and,
occasionally, by β-1,3-glucan bonds (Zhang et al. 2019). These branches consist of
five-carbon sugars (i.e. xylose, rhamnose, and arabinose), six-carbon sugars
(i.e. glucose, mannose, and galactose), and uronic acids. Generally, xylose is the
dominant sugar for hardwoods and agricultural residues, while mannose prevails in
softwoods (Baruah et al. 2018; Singhvi and Gokhale 2019). Contrary to cellulose,
the polymers present in hemicelluloses are easily degradable, due to the amorphous
(non-crystalline) structure and the lower degree of polymerisation. Hemicellulose,
together with lignin, represents a barrier around the cellulose (Fig. 4.1), reducing the
access of cellulases enzymes (Baruah et al. 2018; Zoghlami and Paës 2019). The
most common pretreatments for hemicellulose hydrolysis are dilute acid or alkaline
compounds, steam explosion or enzymes (Zoghlami and Paës 2019).
4.2.3 Lignin
Lignin is the third most abundant polymer in nature, after cellulose and hemicellu-
loses, and generally represents 10–25% of the total feedstock dry matter (Balat
2011). Lignin is an aromatic, complex, three-dimensional cross-linked polymer
synthesized from phenylpropanoid precursors (Fig. 4.1). The lignin structure con-
sists of phenyl propane structural units, liked by aryl ether linkages. The structural
variableness is given by the substitution of the methoxyl groups present in the
aromatic rings (Baruah et al. 2018). The lignin content ranges from 10 to 20% in
various LMs, such as straws, hulls, bagasse, and stalks, and can increase to 30–40%
for nut shells and pinewood (Ponnusamy et al. 2019). The three most common
92 A. Oliva et al.
monomers present in lignin are sinapyl alcohol, coniferyl alcohol, and p-coumaryl
alcohol, corresponding to the three main structural units, which are syringyl,
guaiacyl, and hydroxyphenyl, respectively (Ralph et al. 2019).
Lignin protects the plants from microbial attack and oxidation and gives rigidity
and impermeability to their structure (Ponnusamy et al. 2019). The presence of
lignin is one of the main drawbacks of using LMs in fermentation and AD, as it
makes lignocelluloses resistant to chemical and biological degradation by reducing
the hydrolysis rate (Reddy and Yang 2005). Organic solvents, thermal, and fungal
pretreatment are generally suggested for efficient lignin removal (Amin et al. 2017;
Singhvi and Gokhale 2019).
The biodegradation of LMs is influenced by four main factors, i.e. (1) accessible
surface, (2) crystallinity and degree of polymerization of cellulose, as well as the
(3) lignin and (4) hemicellulose content (Mancini et al. 2016a). Cellulose accessi-
bility is a key factor in the bioconversion of LMs to fermentable sugars. Thus, the
contact between cellulose and cellulase is one of the most critical factors affecting
the enzymatic hydrolysis yield and rate. The contact area depends on biomass
porosity and particle size (Meng and Ragauskas 2014; Xu et al. 2019). Cellulase
accessibility to cellulose mainly depends on porosity, rather than the external surface
of the substrate (Siqueira et al. 2017). Over 90% of the substrate enzymatic digest-
ibility depends on the substrate porosity (Wang et al. 2012). In AD, a limited access
to cellulose results in a scarce contact between biomass and hydrolytic bacteria,
which reduces the release of fermentable sugars for the subsequent degradation
steps. The accessible surface increases along the AD process proportionally with
the degradation of the cell wall components (Xu et al. 2019). The microorganisms-
substrate contact controls the hydrolysis efficiency, especially during the first days of
AD. On the contrary, other factors prevail later, such as the compact structure and the
degree of crystallinity of the remaining cellulose (Oliva et al. 2021; Xu et al. 2019).
The ordered structure and high crystallinity of cellulose are the main deterrents to
convert it to biofuels. Nevertheless, amorphous regions, which are more accessible
to enzymatic attack, are randomly present in the cellulose structure (Zoghlami and
Paës 2019). The cellulose becomes more accessible by decreasing the crystallinity
degree, enhancing the biofuel production from cellulose-rich materials (Jeihanipour
et al. 2011). As described in Sect. 4.2.1, cellulose is a linear homopolymer composed
of microfibrils, joint together to form fibrils and finally fibers. The degree of
polymerization refers to the average length of the polysaccharide chains. An increase
in the degree of polymerization is reflected in a higher density and tensile strength of
the cellulosic component of LMs, making cellulose hydrolysis more difficult (Hallac
and Ragauskas 2011; Mattonai et al. 2018).
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 93
The use of LMs for methane production is limited due to their resistance to the
enzymatic attack (Mahmood et al. 2019). Therefore, many studies have focused on
developing cost-effective pretreatments to reduce the recalcitrance of LMs
(Table 4.2) (Lee et al. 2021; Matsakas et al. 2020; Oliva et al. 2021). Pretreatments
aim to increase the efficacy of lignocellulose hydrolysis by improving the accessi-
bility to cellulose. This scope can be achieved by removing or altering the lignin and
hemicellulose fraction of LMs (Fig. 4.2) (Ali et al. 2020a; Haldar and Purkait 2021).
To be efficient and economically advantageous, pretreatment methods should
meet the following features: (1) high recovery of carbohydrates, (2) high digestibility
of the cellulose in the subsequent enzymatic hydrolysis, (3) high solid concentrations
as well as a high concentration of free sugars in the liquid fraction, (4) no destruction
of hemicelluloses and cellulose, (5) no formation of possible inhibitors for hydrolytic
enzymes and fermenting microorganisms, (6) cost-effectiveness and low consump-
tion of chemicals, (7) low generation of residues, and (8) low capital and operational
costs (Kumari and Singh 2018; Taherzadeh and Karimi 2008).
Pretreatment techniques can be differently classified. A first classification con-
cerns the pH maintained during the process, with pretreatments grouped in acidic,
neutral and alkaline (Ravindran and Jaiswal 2016). Nevertheless, the most common
classification categorises pretreatments into physical, chemical, physicochemical
and biological (Table 4.2) (Singh et al. 2015).
94 A. Oliva et al.
Table 4.2 Effectiveness of different pretreatment methods on the enhancement of the methane
potential of lignocellulosic materials
Category Pretreatment Substrate Δ CH4 Reference
Physical Ball milling Wheat straw +49% Dell’Omo and
Spena (2020)
Microwave Energy crop (Sida +39% Zieliński et al.
hermaphrodita) (2019)
Extrusion Rice straw +72% (Chen et al. 2014)
Chemical Alkaline (NaOH) Wheat straw +15% Mancini et al.
(2018a)
Acid (H2SO4) Cassava residues +57% Zhang et al.
(2011)
Ionic liquid Rice straw +137% Gao et al. (2013)
([C4mim]Cl/DMSO)
Physicochemical Steam explosion Rubber wood waste +670% Eom et al. (2019)
Liquid hot water Sunflower residues +173% Lee and Park
(2020)
Biological Fungal (white-rot) Rice straw +114% Kainthola et al.
(2019b)
Microbial consortia Napier grass +49% Wen et al. (2015)
(WSD-5)
Microbial consortia Catalpa sawdust +76% Ali et al. (2020b)
(CS-5)
The most common physical pretreatments include different milling (e.g. ball, col-
loid, vibro-energy, roller, and hammer), extrusion and irradiation, with the primary
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 95
objective of increasing the accessible surface area and decreasing cellulose crystal-
linity and degree of polymerisation (Amin et al. 2017).
Milling is a size reduction technique employed to increase the surface/volume
ratio and alter the structure and the degree of crystallinity of LMs, making pretreated
substrates more amenable to cellulase attack (Ravindran and Jaiswal 2016).
Irradiation pretreatments involve gamma rays, electron beam, ultrasounds (see
Sect. 4.7) and microwaves, intending to improve the enzymatic hydrolysis of
lignocelluloses (Taherzadeh and Karimi 2008). The pretreatment effectiveness is
proportional to the lignin content, resulting in a lower efficiency on less recalcitrant
(i.e. low lignin) substrates (Keikhosor et al. 2013).
Extrusion pretreatment relies on the spinning of a single or twin screw into a
temperature-controlled barrel. The mechanical action causes strong shearing forces
between the substrate, the screw, and the barrel, locally increasing pressure and
temperature. Apart from the particle size reduction, those forces alter also the
biomass structure and change the crystallinity of the cellulosic component of the
biomass (Duque et al. 2017; Zheng and Rehmann 2014).
Chemical pretreatments act directly on the main components of the biomass, remov-
ing lignin and hemicellulose or decreasing the crystallinity degree of the cellulose
(Ponnusamy et al. 2019). The chemical agents employed are divided into four main
categories: alkali, acids, salts, and organic solvents.
Alkaline pretreatment involves base solutions, e.g. NaOH, KOH, and Ca(OH)2,
aiming to enhance the digestibility of LMs. The main effects on LMs are lignin and
hemicellulose removal, an increase in porosity and the reduction of polymerisation
and crystallinity degree of the cellulose (Tu and Hallett 2019). Alkaline pretreat-
ments occur at mild conditions, i.e. ambient pressure and temperature, but generally
last over 24 h (Amin et al. 2017). NaOH and KOH are the most employed basic
solutions for alkaline pretreatment, having particular effectiveness for lignin removal
from low lignin content LMs (Baruah et al. 2018). Nevertheless, the recycling of
alkaline solutions is challenging, and Na+ and K+ ions can inhibit the subsequent AD
process (Bianco et al. 2021b). On the other hand, a Ca(OH)2 solution can be more
easily recovered but is less effective than NaOH and KOH based solutions (Amin
et al. 2017).
Acid pretreatment is performed with diluted or concentrated solutions. Organic
acids, such as formic acid, as well as inorganic acids, i.e. sulfuric, nitric, phosphoric,
and hydrochloric, are widely employed (Baruah et al. 2018). Dilute acid
pretreatment (0.1–5%) aims to remove the hemicellulosic component of the biomass
and is effective at high temperatures (100–250 C). On the other hand, acids
concentrated at 30–70% hydrolise both cellulose and hemicellulose and require
temperatures below 100 C (Solarte-Toro et al. 2019). Acid solutions are unable to
dissolve lignin but alter the cellulose-hemicellulose-lignin linkages, which increases
96 A. Oliva et al.
the biodegradability of the solid residues (Amin et al. 2017). Nevertheless, the
formation of inhibitory compounds can occur. Phenolic compounds, aldehydes
and furfurals are the most know inhibitory compounds observed after acid pre-
treatments (Ali et al. 2020a). Acid-pretreated solid residues of LMs require abundant
washing to be ready for AD or other biological processes (Rajan and Carrier 2014).
On the other hand, before undergoing fermentation processes, the pH of the hydro-
lysate has to be neutralised with alkaline solutions (Gonzales et al. 2017), which
increases the overall process costs (Castilla-Archilla et al. 2021) and can create
further inhibition (Bianco et al. 2021b).
Ionic liquids and NMMO (see Sect. 4.6) act on the cellulosic component of LMs
(Halder et al. 2019; Mancini et al. 2016b). Ionic liquids are salts in the liquid state at
room temperature in which isolated ions and cations interact by Coulomb forces.
NMMO is a zwitterion containing localised positive and negative charges in a single
molecule (Böhmdorfer et al. 2017). Ionic liquids and NMMO dissolve cellulose,
which can then be regenerated using an anti-solvent (Mancini et al. 2016a). The
regenerated cellulose shows a lower crystallinity, which is a critical factor for the
bioconversion of LMs (Xu et al. 2019). NMMO is effective at different concentra-
tions in aqueous solutions (Wikandari et al. 2016), the mechanisms of which are
thoroughly discussed in Sect. 4.6.
Organosolv pretreatment (see Sect. 4.5) involves organic solvents such as etha-
nol, methanol, and acetic acid, heated to high temperature to remove lignin and
reduce the recalcitrance of LMs. Lignin is a valuable product and can be recovered at
good purity levels after organosolv pretreatment (Ferreira and Taherzadeh 2020).
Liquid hot water pretreatment, similarly to steam explosion, requires high tem-
perature and pressure to remove hemicellulose and disrupt lignin bonds making the
remaining cellulose more available for AD (Hashemi et al. 2019b; Qiao et al. 2011).
Contrary to steam explosion pretreatment, water remains in the liquid state, and
pressure allows maintaining this status at high temperatures (Ruiz et al. 2020). The
hemicellulose sugars, together with other hydrolysable compounds, are hydrolysed,
making liquid hot water pretreatment ideal for LMs rich in non-structural sugars
(Ravindran and Jaiswal 2016). pH has to be controlled to avoid the formation of
inhibitory compounds (Yang et al. 2018). Liquid hot water pretreatment does not
need particle size comminution but demands a large amount of water (Baruah et al.
2018).
pretreatment are still attempting to meet a cost-effective balance beyond the labora-
tory scale (Koupaie et al. 2019; Wen et al. 2015).
Table 4.3 Effectiveness of organosolv pretreatment for lignin removal (Δlignin) and increment of
the methane potential (ΔCH4) of different lignocellulosic materials
Optimal pretreatment Pretreatment
Substrate condition effectivenessa Reference
Rice straw 50% EtOH, 180 C, 1 h Δlignin: 18% Mancini et al. (2018b)
ΔCH4: +41%
Rice straw 75% EtOH, 150 C, 1 h, Δlignin: 22% Mirmohamadsadeghi
catalyst ΔCH4: +32% et al. (2014)
Wheat straw 50% EtOH, 180 C, 1 h Δlignin: 14% Mancini et al. (2018a)
ΔCH4: +15%
Sugarcane 25% EtOH +10% Ammonia, Δlignin: 49% Hashemi et al. (2019a)
bagasse 70 C, 12 h ΔCH4: +135%
Sweet sorghum 50% PrOH, 160 C, 0.5 h, Δlignin: 25% Ostovareh et al. (2015)
stalks catalyst ΔCH4: +107%
Sunflower 50% EtOH, 160 C, 0.5 h Δlignin: 26% Hesami et al. (2015)
stalks ΔCH4: +124%
Hazelnut skin 50% MeOH, 130 C, 1 h, Δlignin: 9% Oliva et al. (2021)
catalyst ΔCH4: +1700%
Hazelnut skin 50% EtOH, 180 C, 1 h Δlignin: N.O. Mancini et al. (2018b)
ΔCH4: +10%
Almond shell 50% MeOH, 200 C, 1 h Δlignin: N.O. Oliva et al. (2021)
ΔCH4: +7%
Cocoa bean 50% EtOH, 180 C, 1 h Δlignin: N.O. Mancini et al. (2018b)
shell ΔCH4: N.O.
Spent coffee 50% MeOH, 200 C, 1 h, Δlignin: N.O. Oliva et al. (2021)
grounds catalyst ΔCH4: +10%
Forest residues 50% MeOH, 190 C, 1 h, Δlignin: 4% Kabir et al. (2015)
catalyst ΔCH4: +320%
Elmwood 75% EtOH, 180 C, 1 h, Δlignin: 27% Mirmohamadsadeghi
catalyst ΔCH4: +73% et al. (2014)
Pinewood 75% EtOH, 150 C, 0.5 h, Δlignin: N.O. Mirmohamadsadeghi
catalyst ΔCH4: +84% et al. (2014)
Rubber wood 75% EtOH, 210 C, 0.5 h Δlignin: 74% Tongbuekeaw et al.
waste ΔCH4: +179% (2020)
a
N.O. means that no significant effect on the specific parameter was observed
swelling of the cellulose fibers by creating balloons, in which the cellulose starts to
dissolve. A further increase of the water content (25–30%) reduces the cellulose
dissolution and mainly results in the swelling of the cellulose fibers. An NMMO
content lower than 65% proportionally decreases cellulose swelling, with no disso-
lution observed (Cuissinat and Navard 2006). The ability of NMMO in dissolving
cellulose is attributed to its chemical structure, presenting weak polar N-O bonds
with a negative charge on oxygen and positively charged on nitrogen. NMMO tends
to form new hydrogen bonds with both water and cellulose, preferentially with
cellulose until the water content of the aqueous solution is below 17% (Mancini
et al. 2016a; Wikandari et al. 2016).
The interest in NMMO started in the early twentieth century with applications in
the textile industry and culminating with its usage in the Lyocell process (Sayyed
et al. 2019). Recently, the effectiveness of NMMO on cellulose has allowed its use
as a pretreatment for LMs (Table 4.4) (Khoshnevisan et al. 2016; Shafiei et al. 2011;
Sołowski et al. 2020). The dissolution mode (85% NMMO) is recommended for
ethanol production, while a lower NMMO concentration results in a better improve-
ment in terms of methane production (Jeihanipour et al. 2010). The dissolution mode
foresees the complete solubilization and subsequent regeneration of the cellulosic
component of the biomass. The rate of cellulose dissolution is inversely related to the
cellulose concentration and degree of polymerization (Wikandari et al. 2016). The
regeneration of cellulose occurs by adding an anti-solvent, such as boiling water
(Cuissinat and Navard 2006). The regenerated cellulose shows a lower total crystal-
linity index (TCI) and lateral order index (LOI), which facilitate the microbial attack.
On the other hand, NMMO pretreatment in swelling/ballooning mode allows a
higher increase in porosity, with a lower decrease of the crystallinity indexes
(Jeihanipour et al. 2010).
TCI and LOI of cellulose-based materials can be estimated by Fourier-transform
infrared spectroscopy. TCI is expressed as the ratio of the absorbance value at 1375
and 2902 cm1 and is proportional to the entire crystallinity of the sample. LOI is
representative of the ratio between cellulose I and cellulose II, and is calculated as
the infrared spectral ratio 1420/893 cm1. The LOI increases with the crystallinity of
cellulose I and decreases with the increasing crystallinity of cellulose II (Carrillo
et al. 2004; Nelson and O’Connor 1964a, b).
reducing the crystallinity of LMs (Jeihanipour et al. 2010). NMMO pretreatment can
be operated in relatively mild conditions, with temperatures ranging from 90 to
130 C and pretreatment times between 20 min and 30 h (Mancini et al. 2016a).
The pretreatment feasibility strongly depends on solvent recovery and recycling.
The solvent recovery takes place by treating the liquid stream of the NMMO
pretreatment with ion-exchange resins to remove contaminants and subsequent
evaporation of the water, obtaining the monohydrate form of NMMO (Satari et al.
2019). NMMO pretreatment non-recyclable liquid waste streams can be treated with
ozone, with ozonation products being easily biodegradable at neutral pH (Stockinger
et al. 1996). The loss of NMMO during recovery is <2% (Sari and Budiyono 2014).
Nevertheless, the use of NMMO as a pretreatment for LMs on a large scale is still
limited because of the considerable amount of water required for the washing step
(Mancini et al. 2016a).
The efficacy of recovered NMMO depends on the chemical composition of the
LMs. In particular, the pretreatment of lignin-rich LMs reduces the efficiency of
recovered NMMO (Millati et al. 2020), most likely due to negative side reactions
and release of by-products such as tannins, resin acids, and phenolic compounds
(Kabir et al. 2014). The NMMO action during pretreatment does not produce furans,
reducing the risk of inhibition in the subsequent biofuel production processes
(Wikandari et al. 2016). However, leftover NMMO after pretreatment can inhibit
the AD process, even at low (0.5–1%) concentrations (Millati et al. 2020).
Recent studies show the effects of 85% NMMO pretreatment at 120 C on rice straw,
wheat straw, and hazelnut skin (Mancini et al. 2016b, 2018a). The pretreatment is
particularly effective on rice straw, increasing the methane production by 82%, even
though no significant effect on the LOI was observed. On the other hand, NMMO
pretreatment significantly reduces the crystallinity index of pretreated hazelnut skin,
resulting in a higher methane production during the first days of AD (Mancini et al.
2016b). Similarly to hazelnut skin, NMMO pretreatment increases the porosity of
wheat straw, enhancing the specific rate constant Rm from 21 to 32 mL CH4/g VS/d
(Mancini et al. 2018a).
Teghammar et al. (2012) investigated the effect of 85% NMMO pretreatment on
spruce wood, rice straw and triticale straw, observing that only for rice straw a longer
pretreatment time (i.e. 15 h) leads to an inhibition of the AD process. On the other
hand, 15 h NMMO pretreatment is particularly efficient for spruce wood and triticale
straw, especially when no comminution is performed. The negative effect of a longer
pretreatment time is due to the loss of glucan and xylan during pretreatment
(Teghammar et al. 2012). On the contrary, all investigated pretreatment times
(i.e. 1, 3, and 15 h) can positively affect the AD of pinewood, confirming that a
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 105
longer time is required when larger particle size substrates undergo NMMO
pretreatment (Shafiei et al. 2014). The dissolution mode pretreatment (i.e. 85%
NMMO) increases the porosity and reduces the crystallinity of birch wood after
3 h pretreatment, enhancing the methane potential by 47% (Goshadrou et al. 2013).
To the authors’ knowledge, only Purwandari et al. (2013) studied the effect of the
NMMO concentration on pretreatment of LMs, although it is reported to be a key
factor for cellulose hydrolysis and subsequent conversion to methane (Jeihanipour
et al. 2010). The authors investigated 73, 79, and 85% NMMO pretreatment at
90 and 120 C. A longer pretreatment time at 90 C positively affects the subsequent
AD of oil palm empty fruit bunches. In contrast, increasing the pretreatment
temperature to 120 C, the optimal pretreatment time is 3 h, and the significance
of the NMMO concentration is attenuated (Purwandari et al. 2013).
P∙t
Pd ¼
V ∙ TS0
where P is the ultrasonic power (W), t is the pretreatment time (t), V is the volume
(L) of the slurry undergoing ultra-sonication, and TS0 is the initial total solid content
(kg) of the slurry.
Apart from the characteristics of the ultrasonic wave, the operating conditions
also impact the cavitation phenomenon. Temperature has a contrasting effect during
ultrasound pretreatment. High temperatures break down the interactions between the
solute and matrix, such as hydrogen bonds, Van der Waals forces, and dipole
attractions. Nevertheless, the vapour pressure increases with the temperature of the
solvent, with the solvent vapour filling the cavitation bubbles, thus reducing their
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 107
Fig. 4.4 Mechanisms of ultrasound pretreatment: (a) alternation of compression and rarefaction
zones in the liquid media, (b) effect of ultrasonic waves on gas bubbles, (c) effect of ultrasounds on
lignocellulosic structure
power of collapse (Bussemaker and Zhang 2013). While water is the most employed
medium for ultrasound pretreatment, other liquids with a lower polarity (e.g. organic
liquids) are also expected to be efficient. On the other hand, viscous liquids (e.g. oils)
make cavitation harder (Santos et al. 2009).
Cavitation has both thermal and physical effects on the solid particles present in the
media, contributing to the fractionation of the lignocellulosic structure and lysing of
the membranes and cell walls of the LMs (Fig. 4.4c) (Rehman et al. 2013; Santos
et al. 2009). Ultrasound pretreatment reduces the crystallinity and degree of poly-
merisation of cellulose, enhancing its hydrolysis and the solubilisation of the overall
organic matter. Besides, ultrasound pretreatment modifies the surface of LMs and
disrupts the lignin linkages, separating the lignin fraction from the cellulose and
hemicellulose sugars (Bundhoo and Mohee 2018; Bussemaker and Zhang 2013;
Hassan et al. 2018).
The main concern on ultrasounds techniques regards cost end energy aspects.
Although some researchers reported the unfeasibility of ultrasound pretreatment on a
laboratory scale (Bundhoo and Mohee 2018), other studies show its energetic
convenience when employing commercial technologies on a larger scale (Cano
et al. 2015). Apart from the energy and economic aspect, the possible degradation
of cellulose and hemicellulose sugars should be considered when performing ultra-
sound pretreatment for LMs valorisation. Nevertheless, this aspect highly depends
on the lignocellulosic substrate (Bussemaker and Zhang 2013).
108 A. Oliva et al.
Several authors report that ultrasounds can reduce the particle size and increase the
surface area of sewage sludge (Bougrier et al. 2006; Cho et al. 2013; Chu et al. 2001;
Na et al. 2007) and manures (Elbeshbishy et al. 2011). Recently, the same effect has
been observed on LMs, showing that the pretreatment duration is proportional to the
substrate disruption. Scanning electron microscopic images show that the damages
to the external surface of wheat straw and maize straw increase with the pretreatment
time. Therefore, ultrasound pretreatment improves the vulnerability of LMs, thus
increasing the available contact surface for the subsequent AD process (Zou et al.
2016a, b).
Ultrasound pretreatment also acts on the chemical composition of LMs. The
cavitation phenomena lead to the formation of radicals that contribute to increased
oxidative stress during pretreatment (Santos et al. 2009). This aspect, together with
high pressure and temperature, reduces the lignin content of LMs, generally leaving
the cellulosic component of the biomass unaffected. On the other hand, the removal
of hemicellulose sugars is observed, most likely due to their amorphous structure and
weak linkages (Perrone et al. 2016). Therefore, ultrasonic waves enable to obtain
cellulose-rich materials, increasing the digestability of LMs during AD. Imam and
Capareda (2012) investigated the effect of hot water and ultrasound-assisted hot
water pretreatment on sweet sorghum residues. Hot water pretreatment alone is not
effective on lignin removal. On the other hand, the ultrasonic implementation
reduced the lignin content by 48%, compared to the untreated sweet sorghum.
Similarly, ultrasound pretreatment of sugarcane bagasse reduced the lignin content
from 27 to 21%, with minor effects on the cellulose and hemicellulose content
(Ramadoss and Muthukumar 2014). Ultrasound-ammonia pretreatment removed
70% of the lignin from sugarcane bagasse (Velmurugan and Incharoensakdi 2016).
While ultrasound pretreatment generally does not change the cellulose content,
ultrasonic waves may affect the cellulosic hydrogen bonds (Nakashima et al. 2016).
In particular, ultrasounds reduce the crystallinity of the cellulosic part of LMs by
rearranging the cellulose structure, leaving a more amorphous polymer, vulnerable
to enzymatic attack (Bussemaker and Zhang 2013). On the other hand, some studies
report an increase in the crystallinity index after ultrasound pretreatment. This is
likely a consequence of the high lignin removal, resulting in a higher cellulose
content remaining in the solid phase after the pretreatment.
The methane potential increases as a consequence of the changed structural and
chemical characteristics of the ultrasound pretreated LMs (Subhedar and Gogate
2014). Qi et al. (2021) investigated the effectiveness of ultrasound pretreatment on
the co-digestion of cannabis straw and municipal wastewater, obtaining a 77%
increment of the methane production after 30 min pretreatment at 100 W. Zou
et al. (2016a, b) observed a positive effect on methane production from wheat and
maize straw co-digested with dairy manure after ultrasound pretreatment.
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 109
References
Ali N, Zhang Q, Liu ZY, Li FL, Lu M, Fang XC (2020a) Emerging technologies for the
pretreatment of lignocellulosic materials for bio-based products. Appl Microbiol Biotechnol
104:455–473. https://doi.org/10.1007/s00253-019-10158-w
Ali SS, Kornaros M, Manni A, Sun J, El-Shanshoury AERR, Kenawy ER, Khalil MA (2020b)
Enhanced anaerobic digestion performance by two artificially constructed microbial consortia
capable of woody biomass degradation and chlorophenols detoxification. J Hazard Mater 389.
https://doi.org/10.1016/j.jhazmat.2020.122076
Alonso-Fariñas B, Oliva A, Rodríguez-Galán M, Esposito G, García-Martín JF, Rodríguez-
Gutiérrez G, Serrano A, Fermoso FG (2020) Environmental assessment of olive mill solid
waste valorization via anaerobic digestion versus olive pomace oil extraction. Processes
8. https://doi.org/10.3390/PR8050626
Amin FR, Khalid H, Zhang H, Rahman S, Zhang R, Liu G, Chen C (2017) Pretreatment methods of
lignocellulosic biomass for anaerobic digestion. AMB Express 7. https://doi.org/10.1186/
s13568-017-0375-4
Arkell A, Olsson J, Wallberg O (2014) Process performance in lignin separation from softwood
black liquor by membrane filtration. Chem Eng Res Des 92:1792–1800. https://doi.org/10.1016/
j.cherd.2013.12.018
Aslanzadeh S, Berg A, Taherzadeh MJ, Sárvári Horváth I (2014) Biogas production from
N-Methylmorpholine-N-oxide (NMMO) pretreated forest residues. Appl Biochem Biotechnol
172:2998–3008. https://doi.org/10.1007/s12010-014-0747-z
Balat M (2011) Production of bioethanol from lignocellulosic materials via the biochemical
pathway: a review. Energy Convers Manag 52:858–875. https://doi.org/10.1016/j.enconman.
2010.08.013
Barbu MC, Sepperer T, Tudor EM, Petutschnigg A (2020) Walnut and hazelnut shells: untapped
industrial resources and their suitability in lignocellulosic composites. Appl Sci 10:1–11. https://
doi.org/10.3390/APP10186340
Baruah J, Nath BK, Sharma R, Kumar S, Deka RC, Baruah DC, Kalita E (2018) Recent trends in the
pretreatment of lignocellulosic biomass for value-added products. Front Energy Res 6:1–19.
https://doi.org/10.3389/fenrg.2018.00141
Behera S, Arora R, Nandhagopal N, Kumar S (2014) Importance of chemical pretreatment for
bioconversion of lignocellulosic biomass. Renew Sust Energ Rev 36:91–106. https://doi.org/10.
1016/j.rser.2014.04.047
110 A. Oliva et al.
Elbeshbishy E, Aldin S, Hafez H, Nakhla G, Ray M (2011) Impact of ultrasonication of hog manure
on anaerobic digestability. Ultrason Sonochem 18:164–171. https://doi.org/10.1016/j.ultsonch.
2010.04.011
Eom T, Chaiprapat S, Charnnok B (2019) Enhanced enzymatic hydrolysis and methane production
from rubber wood waste using steam explosion. J Environ Manag 235:231–239. https://doi.org/
10.1016/j.jenvman.2019.01.041
Ferreira JA, Taherzadeh MJ (2020) Improving the economy of lignocellulose-based biorefineries
with organosolv pretreatment. Bioresour Technol 299:122695. https://doi.org/10.1016/j.
biortech.2019.122695
Gao J, Chen L, Yuan K, Huang H, Yan Z (2013) Ionic liquid pretreatment to enhance the anaerobic
digestion of lignocellulosic biomass. Bioresour Technol 150:352–358. https://doi.org/10.1016/
j.biortech.2013.10.026
Gomez-Tovar F, Celis LB, Razo-Flores E, Alatriste-Mondragón F (2012) Chemical and enzymatic
sequential pretreatment of oat straw for methane production. Bioresour Technol 116:372–378.
https://doi.org/10.1016/j.biortech.2012.03.109
Gonzales RR, Kumar G, Sivagurunathan P, Kim SH (2017) Enhancement of hydrogen production
by optimization of pH adjustment and separation conditions following dilute acid pretreatment
of lignocellulosic biomass. Int J Hydrogen Energy 42:27502–27511. https://doi.org/10.1016/j.
ijhydene.2017.05.021
Goshadrou A, Karimi K, Taherzadeh MJ (2013) Ethanol and biogas production from birch by
NMMO pretreatment. Biomass Bioenergy 49:95–101. https://doi.org/10.1016/j.biombioe.2012.
12.013
Gür TM (2016) Comprehensive review of methane conversion in solid oxide fuel cells: prospects
for efficient electricity generation from natural gas. Prog Energy Combust Sci 54:1–64. https://
doi.org/10.1016/j.pecs.2015.10.004
Haldar D, Purkait MK (2021) A review on the environment-friendly emerging techniques for
pretreatment of lignocellulosic biomass: mechanistic insight and advancements. Chemosphere
264:1–16. https://doi.org/10.1016/j.chemosphere.2020.128523
Halder P, Kundu S, Patel S, Setiawan A, Atkin R, Parthasarthy R, Paz-Ferreiro J, Surapaneni A,
Shah K (2019) Progress on the pre-treatment of lignocellulosic biomass employing ionic liquids.
Renew Sust Energ Rev 105:268–292. https://doi.org/10.1016/j.rser.2019.01.052
Hallac BB, Ragauskas AJ (2011) Analyzing cellulose degree of polymerization and its relevancy to
cellulosic ethanol. Biofuels Bioprod Biorefin 5:215–225. https://doi.org/10.1002/bbb
Harmsen P, Huijgen W, Bermudez L, Bakker R, López L, Bakker R (2010) Literature review of
physical and chemical pretreatment processes for lignocellulosic biomass. Energy. https://doi.
org/10.1016/j.psep.2011.08.004
Hashemi SS, Karimi K, Majid Karimi A (2019a) Ethanolic ammonia pretreatment for efficient
biogas production from sugarcane bagasse. Fuel 248:196–204. https://doi.org/10.1016/j.fuel.
2019.03.080
Hashemi SS, Karimi K, Mirmohamadsadeghi S (2019b) Hydrothermal pretreatment of safflower
straw to enhance biogas production. Energy 172:545–554. https://doi.org/10.1016/j.energy.
2019.01.149
Hassan M, Umar M, Mamat T, Muhayodin F, Talha Z, Mehryar E, Ahmad F, Ding W, Zhao C
(2017) Methane enhancement through sequential thermochemical and sonication pretreatment
for corn stover with anaerobic sludge. Energy Fuels 31:6145–6153. https://doi.org/10.1021/acs.
energyfuels.7b00478
Hassan SS, Williams GA, Jaiswal AK (2018) Emerging technologies for the pretreatment of
lignocellulosic biomass. Bioresour Technol 262:310–318. https://doi.org/10.1016/j.biortech.
2018.04.099
Hesami SM, Zilouei H, Karimi K, Asadinezhad A (2015) Enhanced biogas production from
sunflower stalks using hydrothermal and organosolv pretreatment. Ind Crop Prod 76:449–
455. https://doi.org/10.1016/j.indcrop.2015.07.018
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 113
Mancini G, Papirio S, Lens PNL, Esposito G (2018a) Increased biogas production from wheat straw
by chemical pretreatments. Renew Energy 119:608–614. https://doi.org/10.1016/j.renene.2017.
12.045
Mancini G, Papirio S, Lens PNL, Esposito G (2018b) Anaerobic digestion of lignocellulosic
materials using ethanol-organosolv pretreatment. Environ Eng Sci 35:953–960. https://doi.
org/10.1089/ees.2018.0042
Mancini G, Papirio S, Riccardelli G, Lens PNL, Esposito G (2018c) Trace elements dosing and
alkaline pretreatment in the anaerobic digestion of rice straw. Bioresour Technol 247:897–903.
https://doi.org/10.1016/j.biortech.2017.10.001
Martín C, Wu G, Wang Z, Stagge S, Jönsson LJ (2018) Formation of microbial inhibitors in steam-
explosion pretreatment of softwood impregnated with sulfuric acid and sulfur dioxide.
Bioresour Technol 262:242–250. https://doi.org/10.1016/j.biortech.2018.04.074
Martins F, Felgueiras C, Smitkova M, Caetano N (2019) Analysis of fossil fuel energy consumption
and environmental impacts in european countries. Energies 12:1–11. https://doi.org/10.3390/
en12060964
Matsakas L, Sarkar O, Jansson S, Rova U, Christakopoulos P (2020) A novel hybrid organosolv-
steam explosion pretreatment and fractionation method delivers solids with superior thermo-
philic digestibility to methane. Bioresour Technol 316:1–10. https://doi.org/10.1016/j.biortech.
2020.123973
Matthews JF, Himmel ME, Crowley MF (2012) Conversion of cellulose Iα to Iβ via a high
temperature intermediate (I-HT) and other cellulose phase transformations. Cellulose 19:297–
306. https://doi.org/10.1007/s10570-011-9608-x
Mattonai M, Pawcenis D, del Seppia S, Łojewska J, Ribechini E (2018) Effect of ball-milling on
crystallinity index, degree of polymerization and thermal stability of cellulose. Bioresour
Technol 270:270–277. https://doi.org/10.1016/j.biortech.2018.09.029
Meng X, Ragauskas AJ (2014) Recent advances in understanding the role of cellulose accessibility
in enzymatic hydrolysis of lignocellulosic substrates. Curr Opin Biotechnol 27:150–158. https://
doi.org/10.1016/j.copbio.2014.01.014
Meng X, Bhagia S, Wang Y, Zhou Y, Pu Y, Dunlap JR, Shuai L, Ragauskas AJ, Yoo CG (2020)
Effects of the advanced organosolv pretreatment strategies on structural properties of woody
biomass. Ind Crop Prod 146:112144. https://doi.org/10.1016/j.indcrop.2020.112144
Millati R, Wikandari R, Ariyanto T, Putri RU, Taherzadeh MJ (2020) Pretreatment technologies for
anaerobic digestion of lignocelluloses and toxic feedstocks. Bioresour Technol 304:122998.
https://doi.org/10.1016/j.biortech.2020.122998
Miller DL, Smith NB, Bailey MR, Czarnota GJ, Hynynen K, Makin IRS (2012) Overview of
therapeutic ultrasound applications and safety considerations. J Ultrasound Med 31:623–634.
https://doi.org/10.7863/jum.2012.31.4.623
Mirmohamadsadeghi S, Karimi K, Zamani A, Amiri H, Horváth IS (2014) Enhanced solid-state
biogas production from lignocellulosic biomass by organosolv pretreatment. Biomed Res Int
2014. https://doi.org/10.1155/2014/350414
Mussatto SI, Fernandes M, Roberto IC (2007) Lignin recovery from brewer’s spent grain black
liquor. Carbohydr Polym 70:218–223. https://doi.org/10.1016/j.carbpol.2007.03.021
Mussatto SI, Carneiro LM, Silva JPA, Roberto IC, Teixeira JA (2011) A study on chemical
constituents and sugars extraction from spent coffee grounds. Carbohydr Polym 83:368–374.
https://doi.org/10.1016/j.carbpol.2010.07.063
Na S, Kim YU, Khim J (2007) Physiochemical properties of digested sewage sludge with ultrasonic
treatment. Ultrason Sonochem 14:281–285. https://doi.org/10.1016/j.ultsonch.2006.06.004
Nadir N, Ismail NL, Hussain AS (2019) Fungal pretreatment of lignocellulosic biomass. In:
Biomass for bioenergy-recent trends and future challenges. IntechOpen, pp 39–57. https://doi.
org/10.5772/intechopen.84239
Nagarajan S, Skillen NC, Irvine JTS, Lawton LA, Robertson PKJ (2017) Cellulose II as bioethanol
feedstock and its advantages over native cellulose. Renew Sust Energ Rev 77:182–192. https://
doi.org/10.1016/j.rser.2017.03.118
116 A. Oliva et al.
Ralph J, Lapierre C, Boerjan W (2019) Lignin structure and its engineering. Curr Opin Biotechnol
56:240–249. https://doi.org/10.1016/j.copbio.2019.02.019
Ramadoss G, Muthukumar K (2014) Ultrasound assisted ammonia pretreatment of sugarcane
bagasse for fermentable sugar production. Biochem Eng J 83:33–41. https://doi.org/10.1016/j.
bej.2013.11.013
Ravindran R, Jaiswal AK (2016) A comprehensive review on pre-treatment strategy for lignocel-
lulosic food industry waste: challenges and opportunities. Bioresour Technol 199:92–102.
https://doi.org/10.1016/j.biortech.2015.07.106
Ravindran R, Jaiswal S, Abu-Ghannam N, Jaiswal AK (2018) A comparative analysis of
pretreatment strategies on the properties and hydrolysis of brewers’ spent grain. Bioresour
Technol 248:272–279. https://doi.org/10.1016/j.biortech.2017.06.039
Reddy N, Yang Y (2005) Biofibers from agricultural byproducts for industrial applications. Trends
Biotechnol 23:22–27. https://doi.org/10.1016/j.tibtech.2004.11.002
Rego F, Soares Dias AP, Casquilho M, Rosa FC, Rodrigues A (2019) Fast determination of
lignocellulosic composition of poplar biomass by thermogravimetry. In: Biomass and
bioenergy. Elsevier, pp 375–380. https://doi.org/10.1016/j.biombioe.2019.01.037
Rehman MSU, Kim I, Chisti Y, Han JI (2013) Use of ultrasound in the production of bioethanol
from lignocellulosic biomass. Energy Educ Sci Technol Part A Energy Sci Res 30:1391–1410
Ruiz HA, Conrad M, Sun SN, Sanchez A, Rocha GJM, Romaní A, Castro E, Torres A, Rodríguez-
Jasso RM, Andrade LP, Smirnova I, Sun RC, Meyer AS (2020) Engineering aspects of
hydrothermal pretreatment: from batch to continuous operation, scale-up and pilot reactor
under biorefinery concept. Bioresour Technol 299:1–16. https://doi.org/10.1016/j.biortech.
2019.122685
Santos HM, Lodeiro C, Capelo-Martínez JL (2009) The power of ultrasound. In: Ultrasound in
chemistry: analytical applications. Wiley, pp 1–16. https://doi.org/10.1002/9783527623501.ch1
Sari FP, Budiyono B (2014) Enhanced biogas production from rice straw with various pretreatment:
a review. Waste Technol 2:17–25. https://doi.org/10.12777/wastech.2.1.17-25
Satari B, Karimi K, Kumar R (2019) Cellulose solvent-based pretreatment for enhanced second-
generation biofuel production: a review. Sustain Energy Fuels. https://doi.org/10.1039/
c8se00287h
Sayyed AJ, Deshmukh NA, Pinjari DV (2019) A critical review of manufacturing processes used in
regenerated cellulosic fibres: viscose, cellulose acetate, cuprammonium, LiCl/DMAc, ionic
liquids, and NMMO based lyocell. Cellulose 26:2913–2940. https://doi.org/10.1007/s10570-
019-02318-y
Sen S, Ganguly S (2017) Opportunities, barriers and issues with renewable energy development – a
discussion. Renew Sust Energ Rev 69:1170–1181. https://doi.org/10.1016/j.rser.2016.09.137
Şenol H (2021) Effects of NaOH, thermal, and combined NaOH-thermal pretreatments on the
biomethane yields from the anaerobic digestion of walnut shells. Environ Sci Pollut Res. https://
doi.org/10.1007/s11356-020-11984-6
Shafiei M, Karimi K, Taherzadeh MJ (2011) Techno-economical study of ethanol and biogas from
spruce wood by NMMO-pretreatment and rapid fermentation and digestion. Bioresour Technol
102:7879–7886. https://doi.org/10.1016/j.biortech.2011.05.071
Shafiei M, Karimi K, Zilouei H, Taherzadeh MJ (2014) Enhanced ethanol and biogas production
from pinewood by NMMO pretreatment and detailed biomass analysis. Biomed Res Int 2014.
https://doi.org/10.1155/2014/469378
Shen J, Yan H, Zhang R, Liu G, Chen C (2018) Characterization and methane production of
different nut residue wastes in anaerobic digestion. Renew Energy 116:835–841. https://doi.org/
10.1016/j.renene.2017.09.018
Singh A, Agrawal M (2008) Acid rain and its ecological consequences. J Environ Biol 29:15–24
Singh J, Suhag M, Dhaka A (2015) Augmented digestion of lignocellulose by steam explosion, acid
and alkaline pretreatment methods: a review. Carbohydr Polym 117:624–631. https://doi.org/
10.1016/j.carbpol.2014.10.012
118 A. Oliva et al.
Singhvi MS, Gokhale DV (2019) Lignocellulosic biomass: hurdles and challenges in its valoriza-
tion. Appl Microbiol Biotechnol 103:9305–9320. https://doi.org/10.1007/s00253-019-10212-7
Siqueira G, Arantes V, Saddler JN, Ferraz A, Milagres AMF (2017) Limitation of cellulose
accessibility and unproductive binding of cellulases by pretreated sugarcane bagasse lignin.
Biotechnol Biofuels 10:1–12. https://doi.org/10.1186/s13068-017-0860-7
Sluiter A, Hames B, Ruiz R, Scarlata C, Sluiter J, Templeton D, Crocker D (2008) Determination of
structural carbohydrates and lignin in biomass. Natl. Renew. Energy Lab. Tech. Rep. NREL/TP-
510-42618
Solarte-Toro JC, Romero-García JM, Martínez-Patiño JC, Ruiz-Ramos E, Castro-Galiano E,
Cardona-Alzate CA (2019) Acid pretreatment of lignocellulosic biomass for energy vectors
production: a review focused on operational conditions and techno-economic assessment for
bioethanol production. Renew Sust Energ Rev 107:587–601. https://doi.org/10.1016/j.rser.
2019.02.024
Sołowski G, Konkol I, Cenian A (2020) Production of hydrogen and methane from lignocellulose
waste by fermentation. A review of chemical pretreatment for enhancing the efficiency of the
digestion process. J Clean Prod 267. https://doi.org/10.1016/j.jclepro.2020.121721
Stockinger H, Kut OM, Heinzle E (1996) Ozonation of wastewater containing
N-methylmorpholine-N-oxide. Water Res 30:1745–1748. https://doi.org/10.1016/0043-1354
(95)00320-7
Su J, Qiu M, Shen F, Qi X (2018) Efficient hydrolysis of cellulose to glucose in water by
agricultural residue-derived solid acid catalyst. Cellulose 25:17–22. https://doi.org/10.1007/
s10570-017-1603-4
Subhedar PB, Gogate PR (2014) Alkaline and ultrasound assisted alkaline pretreatment for inten-
sification of delignification process from sustainable raw-material. Ultrason Sonochem 21:216–
225. https://doi.org/10.1016/j.ultsonch.2013.08.001
Suslick KS (1999) Application of ultrasound to materials chemistry. Annu Rev Mater Sci 29:295–
326. https://doi.org/10.1146/annurev.matsci.29.1.295
Taherzadeh MJ, Karimi K (2008) Pretreatment of lignocellulosic wastes to improve ethanol and
biogas production: a review. Int J Mol Sci 9:1621–1651. https://doi.org/10.3390/ijms9091621
Tajmirriahi M, Karimi K, Kumar R (2021a) Effects of pinewood extractives on bioconversion of
pinewood. Fuel 283:119302. https://doi.org/10.1016/j.fuel.2020.119302
Tajmirriahi M, Momayez F, Karimi K (2021b) The critical impact of rice straw extractives on
biogas and bioethanol production. Bioresour Technol 319:124167. https://doi.org/10.1016/j.
biortech.2020.124167
Tang PL, Abdul PM, Engliman NS, Hassan O (2018) Effects of pretreatment and enzyme cocktail
composition on the sugars production from oil palm empty fruit bunch fiber (OPEFBF).
Cellulose 25:4677–4694. https://doi.org/10.1007/s10570-018-1894-0
Teghammar A, Karimi K, Sárvári Horváth I, Taherzadeh MJ (2012) Enhanced biogas production
from rice straw, triticale straw and softwood spruce by NMMO pretreatment. Biomass
Bioenergy 36:116–120. https://doi.org/10.1016/j.biombioe.2011.10.019
Tongbuekeaw T, Sawangkeaw R, Chaiprapat S, Charnnok B (2020) Conversion of rubber wood
waste to methane by ethanol organosolv pretreatment. Biomass Convers Biorefinery. https://doi.
org/10.1007/s13399-020-00710-4
Tu WC, Hallett JP (2019) Recent advances in the pretreatment of lignocellulosic biomass. Curr
Opin Green Sustain Chem 20:11–17. https://doi.org/10.1016/j.cogsc.2019.07.004
Vasco-Correa J, Khanal S, Manandhar A, Shah A (2018) Anaerobic digestion for bioenergy
production: global status, environmental and techno-economic implications, and government
policies. Bioresour Technol 247:1015–1026. https://doi.org/10.1016/j.biortech.2017.09.004
Velmurugan R, Incharoensakdi A (2016) Proper ultrasound treatment increases ethanol production
from simultaneous saccharification and fermentation of sugarcane bagasse. RSC Adv 6:91409–
91419. https://doi.org/10.1039/c6ra17792a
4 Pretreatment of Lignocellulosic Materials to Enhance their Methane Potential 119
Wada M, Nishiyama Y, Langan P (2006) X-ray structure of ammonia-cellulose I: new insights into
the conversion of cellulose i to cellulose IIIi. Macromolecules 39:2947–2952. https://doi.org/10.
1021/ma060228s
Wang QQ, He Z, Zhu Z, Zhang YHP, Ni Y, Luo XL, Zhu JY (2012) Evaluations of cellulose
accessibilities of lignocelluloses by solute exclusion and protein adsorption techniques.
Biotechnol Bioeng 109:381–389. https://doi.org/10.1002/bit.23330
Wen B, Yuan X, Li QX, Liu J, Ren J, Wang X, Cui Z (2015) Comparison and evaluation of
concurrent saccharification and anaerobic digestion of Napier grass after pretreatment by three
microbial consortia. Bioresour Technol 175:102–111. https://doi.org/10.1016/j.biortech.2014.
10.043
Wikandari R, Millati R, Taherzadeh MJ (2016) Pretreatment of lignocelluloses with solvent
N-methylmorpholine N-oxide. In: Biomass fractionation technologies for a lignocellulosic
feedstock based biorefinery. Elsevier, pp 255–280. https://doi.org/10.1016/B978-0-12-
802323-5.00012-8
Xu N, Zhang W, Ren S, Liu F, Zhao C, Liao H, Xu Z, Huang J, Li Q, Tu Y, Yu B, Wang Y, Jiang J,
Qin J, Peng L (2012) Hemicelluloses negatively affect lignocellulose crystallinity for high
biomass digestibility under NaOH and H2SO4 pretreatments in Miscanthus. Biotechnol Biofuels
5:1–12. https://doi.org/10.1186/1754-6834-5-58
Xu N, Liu S, Xin F, Zhou J, Jia H, Xu J, Jiang M, Dong W (2019) Biomethane production from
lignocellulose: biomass recalcitrance and its impacts on anaerobic digestion. Front Bioeng
Biotechnol 7:1–12. https://doi.org/10.3389/fbioe.2019.00191
Yang X, Li Y, Li S, Oladejo AO, Ruan S, Wang Y, Huang S, Ma H (2017) Effects of ultrasound
pretreatment with different frequencies and working modes on the enzymolysis and the structure
characterization of rice protein. Ultrason Sonochem 38:19–28. https://doi.org/10.1016/j.
ultsonch.2017.02.026
Yang B, Tao L, Wyman CE (2018) Strengths, challenges, and opportunities for hydrothermal
pretreatment in lignocellulosic biorefi neries. Biofuels, Bioprod. Biorefining 1:125–138. https://
doi.org/10.1002/bbb.1825
Yao Y, Bergeron AD, Davaritouchaee M (2018) Methane recovery from anaerobic digestion of
urea-pretreated wheat straw. Renew Energy 115:139–148. https://doi.org/10.1016/j.renene.
2017.08.038
Zhang YHP (2008) Reviving the carbohydrate economy via multi-product lignocellulose
biorefineries. J Ind Microbiol Biotechnol 35:367–375. https://doi.org/10.1007/s10295-007-
0293-6
Zhang Q, Tang L, Zhang J, Mao Z, Jiang L (2011) Optimization of thermal-dilute sulfuric acid
pretreatment for enhancement of methane production from cassava residues. Bioresour Technol
102:3958–3965. https://doi.org/10.1016/j.biortech.2010.12.031
Zhang C, Su H, Baeyens J, Tan T (2014) Reviewing the anaerobic digestion of food waste for
biogas production. Renew Sust Energ Rev 38:383–392. https://doi.org/10.1016/j.rser.2014.
05.038
Zhang Y, He H, Liu Y, Wang Y, Huo F, Fan M, Adidharma H, Li X, Zhang S (2019) Recent
progress in theoretical and computational studies on the utilization of lignocellulosic materials.
Green Chem 21:9–35. https://doi.org/10.1039/c8gc02059k
Zheng J, Rehmann L (2014) Extrusion pretreatment of lignocellulosic biomass: a review. Int J Mol
Sci 15:18967–18984. https://doi.org/10.3390/ijms151018967
Zhong C, Wang C, Wang F, Jia H, Wei P, Zhao Y (2016) Enhanced biogas production from wheat
straw with the application of synergistic microbial consortium pretreatment. RSC Adv 6:60187–
60195. https://doi.org/10.1039/c5ra27393e
Zhou Z, Lei F, Li P, Jiang J (2018) Lignocellulosic biomass to biofuels and biochemicals: a
comprehensive review with a focus on ethanol organosolv pretreatment technology. Biotechnol
Bioeng 115:2683–2702. https://doi.org/10.1002/bit.26788
120 A. Oliva et al.
R. Müller (*)
United Nation Industrial Development Organization (UNIDO), Foz do Iguacu, Brazil
M. A. Vilas Boas · M. S. S. M. Costa · M. B. Remor
Agricultural Engineering Graduate Program, Western Parana State University (UNIOESTE),
Cascavel, Brazil
F. S. Marques · D. G. Martinez
International Center for Renewable Energy – Biogas (CIBiogas), Foz do Iguaçu, Brazil
D. A. Santos
National Institute of Industrial Property (INPI), Curitiba, Brazil
G. Bochmann
Institute for Environmental Biotechnology, University of Natural Resources and Life Science
(BOKU), Vienna, Austria
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 121
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_5
122 R. Müller et al.
concludes that the combination of physical and alkaline pretreatment could provide
degradation of cattle dairy residue and the costs and time in new R&D investigations
can be reduced through the analysis of the technological maturity stage.
5.1 Introduction
Biogas as a renewable energy source can replace conventional fuels to produce heat,
power, and vehicular energy. However, in anaerobic digestion the challenge is to
degrade and convert biomass with high fiber content or lignocellulosic biomass, such
as corn stover, rice straw, sugarcane bagasse, and some animal residue such as
poultry litter and dairy cattle litter into biogas.
A considerable amount of dairy cattle rearing activity occurs in intensive or semi-
intensive confined systems to increase the animal production capacity. In the inten-
sive system, the lactating animal spends practically it is whole time in confinement,
and in the semi-intensive system, the confinement time is 4 h. In both confinement
methods, the environmental conditions in the housing area are controlled for
increasing productivity (Perissinotto et al. 2009).
In intensive confinement, usually the animal is located in a shed. The number of
animals is equivalent to the number of bays, from where it only goes out to feed
itself, spending the rest of the time lying down ruminating. Traditionally, agriculture
and wood residues such as sawdust, corn stover, rice straw, and wood shavings are
used as litter for animals (Natzke et al. 1982). These substrates are rich in organic
matter, which makes it interesting for the production of biogas through anaerobic
digestion after their use is over.
The major challenge in the production of biogas from dairy cattle residues and
cattle litter (manure + wood chips) is that they are difficult to degrade by anaerobic
treatment due to the high lignin content. Wood residues (dust, sawdust, or wood
shavings) and agricultural residues (wheat and corn stover, rice straw, among others)
have high levels of cellulose, hemicellulose, and lignin when compared to animal
waste. For non-residual biomass, the cellulose, hemicellulose and lignin concentra-
tions vary between 51–31%, 17–7%, and 36–30%, respectively (Olsson and Hahn-
Hägerdal 1996).
A possible solution for lignin removal from the dairy cattle residues is to
implement a physical, chemical, or biological pretreatment system or a combination
of these, allowing the anaerobic microorganisms to degrade cellulose and hemicel-
lulose (Mosier 2005).
Data collection and analysis of patent databases from the literature can be used to
set up the suitable pretreatment technology. This strategy has been employed in
many areas of research recently, such as in the field of wind energy (Dubarić et al.
5 Biogas Production from Dairy Cattle Residues: Definition of the. . . 123
Fig. 5.1 Strategy steps for defining the technological pathway through bibliometric analysis
2011), solar energy (Zhao et al. 2015), ethanol production from lignocellulosic
biomasses (Silva Schlittler et al. 2012), or in the field of lignocellulosic fuels
(Toivanen and Novotny 2017), demonstrating that such information can be used to
analyze the evolution and level of technology maturity in those respective fields.
However, there are few similar studies in the field of biogas production, more
particularly on pretreatment technologies for dairy cattle residue for biogas
production.
Academic research is directly related to the technological development (demon-
strated by patent registration) in the field of biogas production (Lora Grando et al.
2017). This is because companies, researchers, and universities around the world
actively document their most relevant discoveries through patents to gain a techno-
logical and commercial advantage before engaging in any scientific publication.
In this context, this work aimed to analyze the stage of maturity and R&D efforts
based on patent data and research articles to suggest which is the most promising and
developed technological route, as demonstrated in Fig. 5.1. The case study used was
the market challenges of dairy cattle litter residues for biogas production.
There are many promising pretreatment technologies available that are carried out
with some level of commercial success for marketable products and easily
implemented on an industrial scale in the biogas production process. The
Table 5.1 compares the main pretreatment processes and their effects on cellulose,
hemicellulose and lignin.
The combination of alkaline and physical pretreatments stands out among the
different technologies for promoting mainly the increase in surface area (ISA),
decrystallization of cellulose (CD) and low formation of inhibitors (IF). These
aspects are very relevant for the anaerobic process. Such pretreatment technologies
124 R. Müller et al.
Table 5.1 Effect of different pretreatments on lignocellulosic substrates (Hendriks and Zeeman
2009)
Effect
Pretreatment ISA CD SH LS ALS IF
Mechanical High High N/D N/D N/D N/D
Irradiation High Low Low N/D N/D Low
Steam explosion High N/D High Low High High
Hot water High No effect High Low Low Low
Acid High N/D High Low High High
Alkaline High N/D Low Moderate High Low
Thermal + acid High No effect High N/D N/D Low
Thermal + alkaline High No effect Low Moderate High Low
Explosion of ammonia High High Low High High Low
Biological High N/D High High High N/D
ISA Increase of surface area, CD Cellulose decrystallization, SH Solubilization of hemicellulose, LS
Solubilization of lignin, ALS alteration of the lignin structure, IF inhibitor formation, N/D Not
determined
A6—(chemical)
A7—(acid)
A8—(alkaline OR NaOH OR “sodium hydride”)
Initially, five search strategies were established (general, physical, biological,
chemical and the combination of these pretreatments). However, since the chemical
route had a higher number of records, two more search strategies were set up for acid
and alkaline chemical pretreatment. The rules for search strategies are listed below:
1. General pretreatment:
Patent: (P1) AND (P2) AND (P3) AND (P4) AND ((P5) OR (P6))
Article: (A1) AND (A2) AND (A3) AND (A4) AND (A6) AND (A7)
2. Physical pretreatment:
Patent: (P1) AND (P2) AND (P3) AND (P4) AND (P5)
Article: (A1) AND (A2) AND (A3) AND (A4)
3. Biological pretreatment:
Patent: (P1) AND (P2) AND (P3) AND (P4) AND (P6) NOT (P7 OR P9
OR P10)
Article: (A1) AND (A2) AND (A3) AND (A5)
4. Chemical pretreatment:
Patent: (P1) AND (P2) AND (P3) AND (P4) AND (P6) NOT (P8)
Article: (A1) AND (A2) AND (A3) AND (A6)
5. Combination pretreatment:
Patent: (P1) AND (P2) AND (P3) AND (P4) AND (P5) AND (P6)
Article: (A1) AND (A2) AND (A3) AND (A4 OR A5 OR A6)
6. Acid chemical pretreatment:
Patent: (P1) AND (P2) AND (P3) AND (P4) AND (P6) AND (P9 OR P10)
NOT (P8)
Article: (A1) AND (A2) AND (A3) AND (A6) AND (A7) NOT (A8)
7. Alkaline chemical pretreatment:
Patent: (P1) AND (P2) AND (P3) AND (P4) AND (P6) AND (P10) NOT (P8)
Article: (A1) AND (A2) AND (A3) AND (A4 OR A5 OR A6) AND
(A8) NOT (A7)
The databases were exported and processed with Microsoft Excel software. A
graphical analysis of the level of technological maturity, as recommended by
Achinas et al. (2017) was performed using the methodology proposed by Wang
et al. (2016). The number of patent applications were considered as an indicator of
technological maturity as shown in Fig. 5.2. The purpose of this comparison was to
find out which technology is placed where on this curve, which contains four stages:
germination, growth, maturity and decline.
The technical and scientific efforts in all pretreatment routes were measured and
compared with articles using the Kiviat diagram, considering the total records of
publication in the period from 1998 to 2018. Finally, a descriptive analysis of the
pretreatment effects on lignocellulosic biomasses similar to the dairy cattle residues
(with litter) was done. The observed parameters were: type of biomass, operating
5 Biogas Production from Dairy Cattle Residues: Definition of the. . . 127
Articles Patents
450
400
Number of records
350
300
250
200
150
100
50
0
1998 2000 2002 2004 2006 2008 2010 2012 2014 2016 2018
Year
Fig. 5.3 Number of articles and patents worldwide in the period 1998–2018
conditions (temperature, exposure time and the dose of the catalyst solution),
increase in methane content and biogas.
There have been 1923 patent applications found worldwide, 50.9% of which were
requested by China, 25.1% by the United States, and only four patents by Brazil.
When correlating the number of articles and patents produced in the last 20 years, it
was noticed that the number of publications of scientific articles (2941 publications)
and patent applications (1923 applications) in the period 1998–2006 did not present
significant changes, varying from 0 to 50 (Fig. 5.3). From 2007 onwards, an increase
128 R. Müller et al.
in the number of records was observed, which was also confirmed by Wang et al.
(2016), who related the effect to the fossil fuel crisis, directly impacting the value of
oil and stimulating investments in other energy sources, particularly on renewable
ones.
Overall, the number of articles published was approximately 35% higher than the
number of patents. Magrí et al. (2017) also observed a higher number of articles,
twice as compared to patent deposits, while investigating the nutrients in the
biodigester effluent. About 70% of patent applications and article publications
occurred in the last 10 years, a convergence of the results was observed by Lora
Grando et al. (2017). It can be assumed that the pretreatment routes are relatively
new and emerging in the race for energy sustainability, especially to replace fossil
fuels in the transportation sector.
Until 2008, the patents number was on average 56% higher than the published
articles, after which the difference decreased, and in 2012 the number of articles
exceeded the number of patents. A possible explanation is the very broad patent
claims where one technology can have several applications in different segments
(Lora Grando et al. 2017). However, as the technologies (patents) are tested in the
laboratory and at the industrial scale with different types of biomass and experimen-
tal conditions, the number of articles is expected to increase more than patents. The
difference between the number of articles and patents supposedly confirms the
technological gap pointed out by Achinas et al. (2017). Conversely, the difference
can represent the level of maturity of research in universities and the lack of
alignment with the market needs.
Another important behavior to be observed is the decrease in the number of
published patents after 2011, which may indicate the transition of maturity level
from the germination to the growth stage. The number of patent documents is an
indicator of the level of technology maturity and is useful to guide economic and
scientific efforts (de Luna and Santos 2017), and can be used as an instrument for
decision-making (Wittfoth et al. 2017). Therefore, the higher is the maturity of a
technological route, the more aligned it will be with the market demand. On the other
hand, the lower the maturity, there is more possibility of a disruptive discovery,
consequently, more effort in R&D and investment.
The technological maturity stage for chemical, physical and biological pretreat-
ments are shown in Figs. 5.4, 5.5, and 5.6, respectively. The chemical and physical
route passed through the first and second stages and reached the third, maturity stage,
where solutions supposedly are more accessible to the market and probably most
competitive. The first stage, germination, was passed in 2011, and the second stage,
growth, in 2016 for both pretreatments.
The biological route showed a conservative behavior in the germination stage,
without any significant leap in the patent number. The analysis shown in Fig. 5.6
reflects the literature review, where biological pretreatment of biomass for biogas
production is not commercially ready due to the low biogas increase or high
operation and implementation costs.
For the combination of pretreatments, the graphical analysis is not representative
due to the low number of patents. This type of combined pretreatment is at the
5 Biogas Production from Dairy Cattle Residues: Definition of the. . . 129
Fig. 5.4 Technological maturity analysis for the chemical pretreatment route
Fig. 5.5 Technological maturity analysis for the physical pretreatment route
Fig. 5.6 Technological maturity analysis for the biological pretreatment route
Fig. 5.7 The ration between patent and article documents number
Kiviat diagram was applied to compare different pretreatments and to define which
one has the higher R&D effort (Fig. 5.7).
Considering the main pretreatments, the chemical route showed the best results,
with 40% and 48% of the scientific (articles) and technical (patents) efforts, respec-
tively, followed by biological and physical methods, which had no significant
difference, and the combination of pretreatments with the lowest effort. These data
5 Biogas Production from Dairy Cattle Residues: Definition of the. . . 131
confirm the hypothesis that chemical pretreatment is the most studied technique and
has been widely used for the degradation of lignocellulosic biomass.
Acid and alkaline chemical pretreatments totaled 21% and 20% of all articles and
patent documents and 51% and 49% in the chemical route, respectively. The total
number of articles or patents can be different from the individual sum because some
documents appeared in more than one search strategy. Other keywords to delimit the
search strategy were tested without a positive effect or with an excessively restrictive
result. For example, this was the case for alkaline pretreatment, where once the
condition “NOT acid” was added; the results were not representative because most
of the patents included the word acid in the claims to describe the process. The same
was not observed in the case of acid pretreatment.
recovery of the catalyst, because it is converted into salts, which are unrecoverable or
incorporated into the biomass during the pretreatment reactions (Mosier 2005).
The alkaline pretreatment route showed better results on substrates with low
lignin content such as grasses, agricultural residues, softwood residues, among
others. In this method, the saponification of the intermolecular ester bonds causes
lignin to rupture and separation of the bonds between hemicellulose and lignin
structures leads to the solubilization of lignin and hemicelulose (Sun et al. 2005;
Eggeman and Elander 2005; Achinas et al. 2017).
In most of the previous studies on alkaline pretreatment, NaOH is used as
catalysts, because of its high efficiency, low production inhibitory compounds and
low cost when compared with the other reagents (Amin et al. 2017). Moreover,
NaOH stands out for being non-toxic or non-inhibitory to the subsequent anaerobic
digestion even if there is any residual NaOH left in the treated biomass (Xie et al.
2011).
With this analysis, it can be assumed that the chemical alkaline pretreatment
routes (particularly using NaOH) are more aligned with the market demand due to
the technological maturity and total R&D effort.
biomass (He et al. 2008), allowing the acidogenic bacteria to ferment the biomass
more easily (Neves et al. 2006; Pei et al. 2014). In addition to a larger surface area,
the alkaline hydrolysis releases soluble compounds that can be easily digested by
methanogenic anaerobic bacteria directly, increasing the methane content in the
biogas (Cheng and Liu 2010).
There are some factors that have significant influence on the alkaline pretreatment
efficiency with NaOH; such as reagent concentration, fiber length and pretreatment
(reaction/exposure) time. Selecting a proper dosage is very important, as this decides
the process efficiency without hampering the process economics. Also, it needs to be
mentioned here that just increasing the dosage does not always results in improved
biogas production. At high NaOH dose (>5–9% by weight) along with lignin the
sugars can also be solubilized which ultimately results in low available substrate for
anaerobic digestion and lower biogas yield. But this phenomenon is highly depended
on the biomass type. Moreover, a very high NaOH concentration can also be
detrimental to the anaerobic digestion process due to the inhibitory function of
high amounts Na+ ions in the solution (Monlau et al. 2013; Pang et al. 2008; Pei
et al. 2014; Xie et al. 2011; Zhu et al. 2010).
The length of biomass fibers can be a limiting factor in the pretreatment process,
requiring the adoption of physical pretreatment to reduce the length/size of the
biomass fibers. For example, during studying the efficiency of alkaline pretreatment
of banana stem for biogas production, Pei et al. (2014) found that the shorter the fiber
length, the higher is the biogas yield and the methane content.
The exposure time of the biomass to the alkaline catalyst (NaOH) also has a direct
relationship with the degradation process but the type of biomass plays a critical role
here. Some authors reported satisfactory results in minutes, hours, and days of
pretreatment, depending on the type of biomass and other experimental conditions
such as temperature (Table 5.2). For example, the residue from a vegetal extraction
process treated with NaOH for 15 min can achieve a 55% increase in biogas
production at the ambient temperature in comparison to the biogas produced from
untreated residue. However, using the same pretreatment conditions, only 25% and
37% increase in biogas production were observed from corn stover and rice straw,
respectively. (Zhao et al. 2014; Zhu et al. 2010).
As already mentioned earlier in the introduction that there are rarely any previous
studies on biogas production from dairy cattle litter let alone on its pretreatment for
process improvement. Hence, here previous studies on similar substrates to dairy
cattle litter are compared for understanding the effect of NaOH pretreatment on
biogas production (Table 5.2). From the NaOH pretreatment of different
lingocellulosic biomass depicted in Table 5.2, it can be concluded that dairy cattle
residue consisting of wood shavings as the litter can be treated using NaOH for
enhancing biogas production. NaOH treatment definitely have a positive effect on
the methane content and/or on the biogas production from different types of ligno-
cellulosic biomass. However, depending on the fiber size and the degree of decom-
position of the wood shavings, the use of a physical pretreatment, combined with
NaOH, can help to achieve a more satisfactory results.
134 R. Müller et al.
Table 5.2 Effect of alkaline pretreatment of different lignocellulosic biomass with NaOH along
with the respective experimental conditions
NaOH Increase in Increase
Temperature Treatment dose methane in biogas
Biomass ( C) time (%) (%) (%) Reference
Woods residues
Spruce shavings 15; 0; 5; 2h 7 84 0 Mirahmadi
Birch shavings 50; 80; 100* 2h 7 74 N/D et al. 2010
15; 0; 5; 50;
80; 100*
Banana stem 30–50 1, 2, 3*, 3 50.9–56.2 37.2 Pei et al.
30–50 4, 5 days 6 85.6 2014
30–50 1, 2, 3*, 9 82.2
30–50 4, 5 days
12 58.3
1, 2, 3*,
4, 5 days
1, 2, 3*,
4, 5 days
Agricultural residues
Rice straw Environment 24 h 5 32.8 25.8 Zhao et al.
Rice Environment 24 h 5 28.4 19.7 2014
straw + sewage
sludge
Residue of veg- 37 15 min 8 0 55 Cheng and
etal extraction Liu 2010
Rice straw 20 21 days 6 N/D 64.5 He et al.
2008
37 3h 1,5 70.9 50.0 Sabeeh
et al. 2020
Corn stover 20 21 days 4 N/D 16.6 Pang et al.
20 21 days 6 71 23.8 2008
20 21 days 8 N/D 23.7
20 21 days N/D
10 22.9
Sunflower 55 3, 6, 12, 0,5 N/D N/D Monlau
stalks 55 24*; 36 h 2 N/D N/D et al. 2013
55 3, 6, 12, 4 36 N/D
55 24*; 36 h N/D
6 N/D
55 3, 6, 12, N/D
10 N/D
24*; 36 h
3, 6, 12,
24*; 36 h
3, 6, 12,
24*; 36 h
Corn stover 20 24 h 1 0 0 (Zhu et al.
20 24 h 2,5 0 0 2010)
20 24 h 5 0 37
20 24 h
7,5 ( ) ( )
Energy crops
Grass silage 1 10
(continued)
5 Biogas Production from Dairy Cattle Residues: Definition of the. . . 135
Acknowledgments To CIBiogas, for the scholarship and the suggestion of the research issue. To
the UNIOESTE, for providing infrastructure and mentoring. To the INPI, for the training to search
and use patent documents. This study was financed in part by the Coordenação de Aperfeiçoamento
de Pessoal de Nível Superior—Brasil (CAPES).
References
Achinas S, Achinas V, Euverink GJW (2017) A technological overview of biogas production from
biowaste. Engineering 3:299–307. https://doi.org/10.1016/J.ENG.2017.03.002
Amin FR, Khalid H, Zhang H et al (2017) Pretreatment methods of lignocellulosic biomass for
anaerobic digestion. AMB Express 7:72. https://doi.org/10.1186/s13568-017-0375-4
Bochmann G, Montgomery LFR (2013) Storage and pretreatment of substrates for biogas
production. In: The biogas handbook. Woodhead Publishing, pp 85–103. https://doi.org/10.
1533/9780857097415.1.85
Chen H, Liu J, Chang X et al (2017) A review on the pretreatment of lignocellulose for high-value
chemicals. Fuel Process Technol 160:196–206. https://doi.org/10.1016/J.FUPROC.2016.
12.007
Cheng XY, Liu CZ (2010) Enhanced biogas production from herbal-extraction process residues by
microwave-assisted alkaline pretreatment. J Chem Technol Biotechnol 85:127–131. https://doi.
org/10.1002/jctb.2278
de Luna AJ, Santos DA (2017) Comparative analysis between academic and patent publications
based on Fenton Technologies among China, Brazil, and the rest of the world. Environ Sci
Pollut Res 24:6106–6113. https://doi.org/10.1007/s11356-016-6570-z
Dubarić E, Giannoccaro D, Bengtsson R, Ackermann T (2011) Patent data as indicators of wind
power technology development. World Pat Inf 33:144–149. https://doi.org/10.1016/J.WPI.
2010.12.005
Eggeman T, Elander RT (2005) Process and economic analysis of pretreatment technologies.
Bioresour Technol 96:2019–2025. https://doi.org/10.1016/j.biortech.2005.01.017
He Y, Pang Y, Liu Y et al (2008) Physicochemical characterization of rice straw pretreated with
sodium hydroxide in the solid state for enhancing biogas production. Energy Fuels 22:2775–
2781. https://doi.org/10.1021/ef8000967
136 R. Müller et al.
Abstract Dairy industries are one of the major food industries which generate a
huge amount of wastewater both in terms of volume and strength. The chemical
composition of dairy wastewater ranges from 1000 to 4500 mg/l for COD, 500 to
3000 mg/l for BOD, and 160 to 800 mg/l for TSS. Such wastewaters, if discharged
into the environment without proper treatment will pose serious detrimental effects
on water, land and air. The major portion of the wastewater generated from dairies is
highly organic which depicts its higher degree of biodegradability. Compared to the
aerobic and physicochemical methods employed for treating dairy wastewaters,
anaerobic treatment methods are highly promising, cost-effective, and provide a
sustainable energy generation option. The present manuscript critically evaluates the
effect of various factors like pH, temperature, organic loading rate, hydraulic
retention time, availability of nutrients, and C/N ratio on the process of anaerobic
digestion. Recent studies and older research are considered for this study which
concludes that these factors have a pertinent influence on the performance of an
anaerobic digester. To check the feasibility of adopting an anaerobic digestion
technology in both small and large-scale dairy units, technical and economic anal-
ysis is a must. This chapter also provides a detailed techno-economic analysis (TEA)
framework for biogas production from dairy wastes.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 141
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_6
142 S. Pilli et al.
6.1 Introduction
“Operation Flood” also known as the “White Revolution” in the year 1970
transformed India from a milk deficient nation to the leading producer of milk in
the world. Since 1997, India is the leading milk-producing country in the world.
Dairy products are considered as nutritious food source. Further, it is also considered
as the means for increasing employment opportunities and thereby boosting the
economy. India is also a leading consumer of dairy products and has both private and
cooperative dairies. In 2019, the milk production was 175 billion litres and by 2020
it is expected to increase by 4%. India contributes to about 22% of the global milk
production. With the growing population, there is a tremendous increase in the
demand for dairy products including milk, curd, cheese, butter, ghee (clarified
butter), ice cream, custard and cream. The dairy industry is one of the food
processing industries which consume large volumes of water in every process
(Singh et al. 2014). During milk processing, around 2% of the processed milk
comes out in the waste, and the wastewater generated is 0.2–10 l/l of milk processed
(Mehrotra et al. 2016).
Milk processing in the dairy industry involves several unit operations such as
receiving, storing, clarification, pasteurization, standardization, homogenization,
deodorization, storage and packaging (Burke et al. 2018; Motarjemi et al. 2014).
For cleaning the various units, different alkaline solutions, detergents, emulsifiers,
and sanitizers are used. In the dairy industry, the wastewater produced is mainly
categorised into three types (Britz et al. 2006):
(i) Processing wastewater: i.e. the wastewater formed while cooling the milk in
coolers and condensers for evaporating milk or whey.
(ii) Cleaning wastewater or sanitary wastewater: The cleaning wastewater evolves
during the washing of process units, storage tanks, and milk cans, in contact
with milk and other milk products. Spillages, losses in packaging, breakdown
of equipment, and by-products (such as whey and brine) spillages/losses are
considered in this category. Cleaning wastewater is around 90% organic as it
comes from milk and milk products manufacturing residues: whey, cream,
cheese pieces, water from clarification and separation (Kolev Slavov 2017).
(iii) Industrial wastewater: includes water used in clean in point (CIP) of equipment
in contact with milk products.
The composition and concentration of wastewater depend on the operation
method, production process and design of the process plant. This wastewater is
highly unstable and biodegradable (Sharma et al. 2008).
Water used in the dairy industry is generally contaminated by milk and various milk
products produced in the processes. The characteristics of the effluents generated in a
6 Anaerobic Digestion of Dairy Industry Wastewater 143
dairy industry are summarised in Table 6.1. Wastewater from the dairy industry
consists of milk solids, detergents, milk wastes and sanitizers. The organic fraction
in the dairy wastewater is because of lactose, glucose, fat and protein present in the
milk products (Hassan and Nelson 2012). The characteristics of dairy wastewater
and its composition depend on the size of the dairy industry and the type of products
produced in the industry. A typical dairy industry processing milk, curd, ice cream,
butter and buttermilk generates wastewater having COD, BOD, TSS in the range of
2500–3000, 1300–1600, 72,000–80,000 mg/l, respectively and a pH of 7.2–7.5
(Raghunath et al. 2016). Whereas, the wastewater from fluid milk processing has a
COD, BOD, and TSS in the range of 500–1300, 950–2400, and 95–450 mg/l,
respectively, and the pH 5–9.5 (Demirel et al. 2005). Further, the COD, BOD, and
TSS of the wastewater produced at the milk reception due to spillage and cleaning of
the milk cans are 2540, 800, and 650 mg/l, respectively and the pH is 7.18
(Janczukowicz et al. 2008). Wastewater from the washing of equipment such as
boilers has a COD, BOD, and TSS concentration of 14,640, 3470 and 3820 mg/l,
respectively and a pH of 10.37 (Janczukowicz et al. 2008).
In many countries, large amounts of dairy products in the form of cheese and
butter are consumed. Cheese whey streams contain protein, lactose and other mineral
elements that can be utilized in the manufacture of different products such as lactic
acid, vitamins, fermented whey drinks, baker’s yeast and antibiotics. However,
recovery of whey or whey protein results in new wastewater streams that also
need to be treated (Malaspina et al. 1995; Oreopoulou and Russ 2007). The waste-
water produced from the cheese whey processing is 60–80 times more concentrated
than domestic sewage. The whey stream has a COD and TSS concentration in the
144 S. Pilli et al.
Due to the high organic content and biodegradability, the dairy wastewater depletes
the dissolved oxygen content of the receiving stream and creates anaerobic condi-
tions resulting in foul odour. A higher concentration of COD and BOD is toxic to
aquatic life. Suspended solids, soluble organics and trace organics in dairy waste-
water lead to eutrophication of the receiving streams. Even if the BOD is low but the
phosphorus and nitrogen content is too high, this could trigger the increased pro-
duction of algae. Wastewater disposed into the streams causes the growth of algae
and bacteria that consume oxygen and suffocate the aquatic life (Deshpande et al.
2012). Lactose, a low molecular weight organic compound present in dairy waste-
water promotes the growth of sewage fungi.
Bacteria convert the nitrogen in protein to the inorganic forms of nitrogen as
nitrate, nitrite, ammonia and ammonium ions. Different forms of nitrogen are
harmful to both humans and livestock. Nitrate ions converted to nitrite ions in the
bloodstream and haemoglobin converted to methemoglobin does not allow oxygen
to be taken up by the blood. Methemoglobinemia is a disease observed in infants due
to the presence of methemoglobin in the bloodstream. To avoid such effects, the
6 Anaerobic Digestion of Dairy Industry Wastewater 145
Fig. 6.1 Pollutants from the milk processing unit and dairy wastewater
The concentration of the dairy wastewater is rich in nutrients and organic matter thus
conventional treatment techniques are not effective in removing them. Mechanical
treatment followed by chemical or biological treatment are generally carried out
(Birwal et al. 2017; Yonar et al. 2018). In a conventional dairy wastewater treatment
system, the preliminary stage includes screens, grit chambers and a skimming tank to
remove floating materials, inorganic soil material and oil and grease from the dairy
effluents respectively. Solids removed from dairy wastewater, such as fats, oils and
greases (FOGs), increase the biodegradability and methane production, but at the
same time decrease the methane production potential by decreasing the BOD
concentration. Figure 6.2 shows the possible treatment ways of dairy wastewater
through mechanical, physico-chemical and biological methods.
Treatment of dairy wastewater through the biological process is promising since
dairy wastewater effluent is mainly organic. Aerobic treatment processes are less
efficient due to the growth of filamentous substances and acidification (high lactose
and low buffer) (Nadais et al. 2010; Prazeres et al. 2012). Bulking and foaming,
additional biomass production, as well as poor activity at low temperature, are the
other disadvantages of aerobic treatment processes (Britz et al. 2006). Anaerobic
treatment followed by aerobic treatment is the most commonly used process for
treating dairy wastewater.
6 Anaerobic Digestion of Dairy Industry Wastewater 147
Anaerobic systems used for the treatment of dairy industry effluent are anaerobic
lagoon, contact digester, upflow anaerobic sludge blanket reactor, stirred tank
reactor, fixed-film digester, anaerobic filter reactor, expanded bed digester, mem-
brane anaerobic digester, separated phase digester and hybrid digesters (Hassan and
148 S. Pilli et al.
Nelson 2012). Anaerobic digestion of dairy wastewater produces organic acids such
as lactic acid, propionic acid and acetic acid, and biogas with hydrogen, methane and
carbon dioxide. In some cases, methanogenesis is inhibited for the production of
other products such as hydrogen, propionic and acetic acid, which is called dark
fermentation (Kasmi 2018).
There are many studies focusing on anaerobic treatment of dairy wastewater with
very high treatment efficiency. The COD removal of various dairy wastewaters and
biogas production presented in the literature is summarised in Table 6.2. Gavala
et al. (1999) stated that an up-flow anaerobic sludge blanket (UASB) reactor under
steady-state conditions removed 90% of COD from the influent having a COD of
2050 mg/l and organic loading rate of 0.031 kg of COD/m3 day. An anaerobic
UASB reactor followed by an aerobic reactor resulted in 90% and 85% removal of
BOD and COD respectively, in a cheese manufacturing industry with a biogas
production rate of 0.40 l biogas/g of COD removed (Malaspina et al. 1995). At an
organic loading rate of 10 g/day, an anaerobic downflow-upflow hybrid reactor
(DUHR) removed 98% COD followed by an anaerobic sludge blanket reactor
(SBR) that achieved 90% removal of both nutrients and COD (Derramadero and
Guyot 1995).
A combination of an anaerobic filter reactor and UASB reactor to treat dairy
wastewater was evaluated by Calli and Yukselen (2002), and it was concluded that
the digester generated 0.354 m3 methane/kg of COD (Calli and Yukselen 2002). The
performance of an anaerobic fixed film reactor treating dairy wastewater at 3 and
2 days HRT was evaluated by Koshta (2010) and observed a BOD removal
efficiency of 87.69% and 89.42%, respectively. The performance of anaerobic
sequence batch reactors (ASBR) was evaluated at an HRT of 6 h and a temperature
of 58 C and observed a removal efficiency of 62% and 75% of COD and BOD5,
respectively (Dugba and Zhang 1999). Further, Nadais et al. (2006) evaluated an
intermittent feeding cycle, i.e. 48 h feed + 48 h without feeding, to the UASB reactor
and observed that intermittent operation has a higher COD removal efficiency (22 g
COD/l/day) compared with continuous feeding (3–6 g COD/l/day). Dębowski et al.
(2018) treated the simulated dairy wastewater in an anaerobic biofilm reactor with
magneto active microporous media. At a suitable organic loading rate ranging from
6 to 8 kg COD/m3 day, the anaerobic reactor removed 80% of chemical oxygen
demand, 70% of organic matter and 80% of phosphorous. The methane production
was in the range of 420.6–557.1 l/day. Anaerobic digestion of the dairy wastewater
as a substrate was performed by Karthiyayini et al. (2017) and observed 36% COD
and 33.5% of volatile solids removal. Similarly, many recent studies suggest that
anaerobic digestion is a more energy-efficient process for dairy wastewater treatment
with increased biogas production (Shete and Shinkar 2017; Meegoda et al. 2018;
Pilli et al. 2020).
The anaerobic digestion of milk fat (triglycerides) in dairy wastewater by micro-
organisms (Clostridia and Micrococci sp.) results in the products such as acetate,
hydrogen and CO2. But, the presence of long-chain fatty acids (LCFAs) in fats, oils
and greases (FOGs) can cause inhibition of methanogenesis and digester instability.
In anaerobic digestion, lactose generates acetate, lactate, formate, propionate and
Table 6.2 Process performance of the anaerobic digestion of dairy wastewater
OLR COD Methane
HRT (kg COD/ Temperature removal production
Dairy waste type Reactor (Days) m3 day) pH ( C) (%) (m3/kg COD) Reference
Ice cream Anaerobic filter 0.5 6 – – 85 0.32–0.34 Ince (1998)
Acid whey CSTR 5 – 6.5 – – 0.3 Saddoud et al.
(2007)
Raw milk Anaerobic filter 0.5 5–6 – – 90 – Omil et al. (2003)
Cheese plant CSTR 5.7 60 7.0 - - 0.28 Hwang (1997)
waste
Cheese whey Downflow fixed-film 6.6 8.3 – 30 76 0.34 Van den Berg and
Kennedy (1983)
Deproteinized Downflow fixed bed 5 2.8 7.25 – – 27 De Haast et al.
whey (1986)
Effluent of inte- UASB 0.2083 8.5 – 30 87 – Ozturk et al.
grated plant (1993)
6 Anaerobic Digestion of Dairy Industry Wastewater
(continued)
Table 6.2 (continued)
150
There are many important process parameters in anaerobic digestion and by opti-
mizing them, an increase in biogas yield and treatment efficiency can be achieved.
The main parameters include pH, temperature, organic loading rate, HRT, nutrients
and mixing (Abdelgadir et al. 2014), which are discussed in detail below.
6.3.1 pH
6.3.2 Temperature
The organic loading rate (OLR) of the anaerobic digester is the total amount of COD
in kg/m3/day of digester volume supplied to the digester. It is the total amount of
organics fed into the digester in a day (Shete and Shinkar 2017). An increase in OLR
beyond the optimum level decreases the methane production in the digester and the
COD removal efficiency from the dairy wastewater. Overloading the digester causes
wastewater to quickly hydrolyze and acidify thus creating over-accumulation of
volatile fatty acids, which has the potential of inhibiting methanogenesis (Meegoda
et al. 2018). A higher organic loading rate often causes inhibition of methanogens in
the digester and causes reactor failure. Some reactors like fixed film, fluidized bed
and expanded bed reactors can withstand high organic loading rates. The desired
OLR of the digesters is the provision of the possible highest substrate concentrations
to methanogens for methane production while compensating the inhibitory com-
pounds (Tabatabaei et al. 2011).
Hydraulic retention time (HRT) is the mean length of time that liquids remain in a
digester. A higher loading rate reduces the hydraulic retention time, which may
affect the efficiency of digestion (Meegoda et al. 2018). Accumulation of volatile
fatty acids increases with a decrease in hydraulic retention time. Long hydraulic
6 Anaerobic Digestion of Dairy Industry Wastewater 153
6.3.5 Nutrients
In the anaerobic digester, along with the substrate and inoculum, (micro-)nutrients
are also essential for anaerobic digestion. Inorganic nutrients such as nickel, iron,
cobalt and other trace minerals are added to the digester to enhance the anaerobic
digestion by enhancing the microbial population growth (Shete and Shinkar 2017).
In case of dairy wastewater, macronutrients such as nitrogen and phosphorous are
already present. Calcium increases the granulation, which helps in retaining slow-
growing methanogens in the digester and improve the treatment efficiency (Hassan
and Nelson 2012; Demirel and Yenigun 2004).
Biomass and various industrial wastes have been used as sources of renewable
energy production through anaerobic digestion for biogas production. The biogas
produced through anaerobic digestion has been used in ranges of 1–6 m3 for
154 S. Pilli et al.
household purposes and more than 1000 m3 for industrial purposes (Deublein and
Steinhauser 2011). Dairy industries have huge potential in producing biogas from
the highly organic effluents it generates. But many dairies see the biogas generation
strategy as a high-risk investment due to the high sensitivity of the process, varia-
tions in feedstock characteristics, segregation of wastes and operational problems
(Schmidt 2014). Many governments have put several policy initiatives and off-take
agreements to solve these issues. For this, at first, it is important to conduct an
integrated analysis to assess the technical and economic feasibility and environmen-
tal impacts of employing anaerobic digestion technologies.
A techno-economic analysis (TEA) indicates the technical and economic analysis
of a process/technology through various simulation approaches through software
like Aspen plus and Intelligen Superpro Designer. The results provide an insight into
the feasibility of a project. Also in certain cases, Life Cycle Assessments (LCA) are
used to compare the impacts of various processes/products quantitatively on the
environment (Li and Khanal 2016). The whole process of anaerobic digestion can be
divided into three areas: collection and transportation of feedstock, pre-treatment and
anaerobic digestion, and post-processing of biogas/co-product handling. Similarly,
in TEA for the preparation of TEA framework, three basic sections can be identified
namely (a) the unit operations involved in choosing, collecting and transporting
feedstock to the digestion site, (b) processing stage of feedstock to produce biogas,
and (c) final stage of upgrading the biogas generated for electricity generation,
heating, liquid methane and house-hold purposes (Rajendran and Murthy 2019).
Large dairy industries can utilise the wastes generated for electricity generation and
heating purposes. by installing an anaerobic digestion plant within the industry,
because the wastes generated may be large enough to generate the energy required
for heating. Mostly large dairies can afford a one-time capital on setting up a
treatment plant within the industrial premises. This will also help in reducing the
collection and transportation cost in transferring wastes from the source to the
digestion site. Hence, the first section in a TEA framework pertaining into operations
involving the transportation of wastes can be neglected and this will help in reducing
the complexity. Figure 6.3 represents a typical TEA framework for biogas produc-
tion from dairy wastes.
The technical goals include maximising dairy waste utilization for AD,
maximising biogas yield, and process stabilisation. The economical objective is to
reduce the costs, i.e. capital, operational and maintenance costs. The whole waste
processing value chain is determined by the TEA framework and through experi-
mentally validated process specifics. The economic analysis involves cost assess-
ments and investment analysis. The techno-economic analysis is conducted to
estimate the cost required and energy generated over the value chain from feedstock
digestion to energy production by various means like combined heat and power
(CHP), electricity, or vehicle fuel. Studies on the use of biogas, as vehicle fuel are
one of the leading technological advancements in the application of biogas.
Table 6.3 shows a summary of AD studies on dairy wastes, which types of wastes
used, OLR, other operational parameters, biogas yield and expenditures encountered
are summarised.
6 Anaerobic Digestion of Dairy Industry Wastewater
Fig. 6.3 TEA framework for biogas production from dairy wastes
155
156
Pilot-scale studies are conducted more in number than field scale/industrial scale
set-ups. Some of the few recent anaerobic digestion studies on dairy wastes includ-
ing their techno-economic analysis are discussed below.
Data regarding electricity, diesel, natural gas and water consumption by the five
dairies are furnished. Heat and electricity that can be derived were calculated by
burning biogas in a combined heat and power unit (CHP) having 50% electricity and
35% heat. Energy consumption of the five dairies is collected. Thermal consumption
rates were high (0.247–0.557 MJ/kg milk) compared to the values presented in the
literature. Among the five dairies, methane yield was high for dairies 2 and
158 S. Pilli et al.
4, i.e. 307.7 Sm3/day and 317.8 Sm3/day, respectively. Based on the values obtained,
an appropriate digester size that can be installed in each dairy-based on cheese yield
was proposed. Results showed that each dairy gets some surplus energy after
covering most of the plant energy needs if they start digesting all the whey produced
by themselves. Thus, the anaerobic digestion of dairy waste/wastewater approach is
a way for sustainable development. Further, the transportation and management
costs are significantly minimised along with renewable energy production and
improved energy balance.
Tan et al. (2021) has conducted anaerobic digestion studies for determining the
suitability of using an attached biofilm reactor with cattle manure as the substrate.
Techno-economic evaluation of the biogas system was also done. The cow manure
collected from a dairy at Ladang (Malaysia) was cultured overnight and then
acclimatised using a sequencing batch reactor for use as an inoculum. The digester
was set to operate at 37 C, HRT of 7 days and pH 7. Granular Activated Carbon
(GAC) was used as a support carrier on which the biofilm layer was formed. The
percentage of GAC used was 25% (v/v) which means, 25% of the total working
volume was attached over the support carrier, which helps in the accumulation of
microorganisms on the biofilm layer formed. The maximum methane production of
934.54 mL/g VS was observed at 14 days HRT. Using the modified Gompertz
equation methane production was estimated to be the maximum on 20 days HRT
period.
An energy economic analysis was conducted for the study. The lifespan of the
digester system is estimated to be 20 years and the rate of electricity is 0.35
Malaysian Ringgit (RM) /kWh. The cost spent on the digester set up was divided
into fixed capital and working capital, where fixed capital is determined based on
some factors (equipment cost, purchase cost, piping and site development cost) and
working capital is 5% of the fixed capital. Also, operational cost is the sum of fixed
operating cost and variable cost. The generated biogas was used for heating and
electricity generation purposes. A combined heat and power (CHP) unit having 48%
thermal efficiency and 38% electrical efficiency was operated using the methane
generated. The revenue obtained from the sale of electricity generated using biogas
was around RM 236.88 and RM 2842.56 in a year. This electricity can be effectively
utilised within the farm which will reduce the electricity bills enormously. However,
this alone cannot help in covering the total operational cost. Thus, fertiliser formed in
6 Anaerobic Digestion of Dairy Industry Wastewater 159
the digester can also be supplied to the farmers which will help in expanding the
revenue. At the initial stages, the cash flow was a little down as the profit was used to
pay off the debts. Later, from 2 to 7 years the revenue from the biogas plant showed
increment.
Fantozzi et al. (2015) studied the potential of energy production through anaerobic
digestion of spoiled milk using a pilot-scale anaerobic plant. Spoiled milk comprises
a major portion of waste generated from the dairy industry. Milk gets spilt out during
many practices like collection, transferring milk from cans, processing units and
packaging. The whole fresh milk that got expired and stored in warehouses are used
for this study. The digestate from a secondary anaerobic digester fed with corn waste
and sorghum silage was used as inoculum. The experimentation involves
biomethanation (BM) tests in bottles and a pilot-scale set up. The substrate to
inoculum ratio of 1:3 fixed from BM tests is used for the pilot-scale study. The
study was carried out at a temperature of 35 C for 40 days. Usage of the optimum
amount of organic matter has avoided the rapid acidification problem and attained a
maximum methane production of 0.362 Nm3/kg VS.
The anaerobic digester plant consists of a CSTR reactor with a pre-tank load, a
cogeneration unit to produce electricity and heat and a biogas storage unit. The
power unit is in the micro-scale range (below 200 Kw). The net electric output power
is around 105 kW and thermal power of 127 kW. The small to medium dairies
located near the plant process around 15,000 tons of milk every year and 10% of the
product normally becomes waste which is utilised for producing biogas. The diges-
tion plant is operated and maintained by a dairy plant which collects raw material
from the three other dairies. Since the dairies are located with a 50 km distance,
transportation cost is not much high. Also, the benefit of supplier companies is their
avoided transportation costs to landfills. The revenue from the incentives was
calculated considering that about 8% of the electric power is consumed for plant
operation. Disposal costs are imposed on other dairies which will also add to the
revenue earned. Around 8% of the electric power generated is used for operating the
plant.
160 S. Pilli et al.
From the bioreactor point of view, the two-stage and high-rate anaerobic digestion
systems for dairy wastewater are attractive for achieving process energy efficiency.
The production of hydrogen from dairy wastewater along with methane could also
be a promising strategy (Demirel and Yenigun 2004; Karadag et al. 2014; Murari
et al. 2019). In addition, highly polluting dairy wastes like whey wastes can be
processed further to derive several value added byproducts (Asunis et al. 2020). The
two-stage anaerobic digester consists of two reactors, an acidogenesis or
hydrogenogenesis reactor and a methanogenesis reactor. In high-rate anaerobic
digesters, the substrate is completely mixed, operated at elevated temperature and
the retention time is lower. Such high-rate systems are more beneficial from an
economic point of view. Further, techno-economic analysis is critical for
establishing the full-scale process. Nutritional, chemical, mechanical and energy
requirements and the innovative design of anaerobic digesters is essential to improve
methane production and to reduce costs. Another strategy is co-digestion which
utilizes multiple substrates for methane production in anaerobic digestion.
Co-digestion of dairy wastewater with other organic feedstocks is advantageous to
enhance methane production and to achieve a waste sustainable process. Demand for
dairy products are increasing worldwide and as a result increase in dairy industries
and the use of various chemicals (hormones, antibiotics and other drugs) are
becoming more prevalent. Such new chemicals in dairy wastewater may seriously
affect the anaerobic digestion process and the treated water quality. Hence, special
care should be taken to develop innovative research ideas to tackle this problem in
order to prevent toxicity in the anaerobic digestion of dairy wastewater and prevent
environmental degradation.
References
Abdelgadir A, Chen X, Liu J, Xie X, Zhang J, Zhang K, Wang H, Liu N (2014) Characteristics,
process parameters, and inner components of anaerobic bioreactors. BioMed Res Int. Article ID
841573. https://doi.org/10.1155/2014/841573
Akbulut A (2012) Techno-economic analysis of electricity and heat generation from farm-scale
biogas plant: Çiçekdağı case study. Energy 44(1):381–390. https://doi.org/10.1016/j.energy.
2012.06.017
Asha B (2014) Influence of pH condition on the performance fixed film fixed bed reactor in dairy
wastewater. I Control Pollution. J Ind Pollut Control 31(1):33–36
Asunis F, De Gioannis G, Dessì P, Isipato M, Lens PN, Muntoni A, Polettini A, Pomi R, Rossi A,
Spiga D (2020) The dairy biorefinery: integrating treatment processes for cheese whey
valorisation. J Environ Manag 276:111240. https://doi.org/10.1016/j.jenvman.2020.111240
Barnett J, Robertson S, Russell J (1998) Environmental issues in dairy processing.
Trans. E. Portfolio. In: Chemical processes in New Zealand. New Zealand Institute of Chem-
istry, Palmerston North, pp 1–16
6 Anaerobic Digestion of Dairy Industry Wastewater 161
Gavala HN, Kopsinis H, Skiadas IV, Stamatelatou K, Lyberatos G (1999) Treatment of dairy
wastewater using an upflow anaerobic sludge blanket reactor. J Agric Eng Res 73(1):59–63
Gotmare M, Dhoble RM, Pittule AP (2011) Biomethanation of dairy wastewater through UASB at
mesophilic temperature range. Int J Adv Eng Sci Technol 8(1):1–9
Hassan AN, Nelson BK (2012) Invited review: anaerobic fermentation of dairy food wastewater. J
Dairy Sci 95(11):6188–6203
Hwang SH (1997) Feasibility assay in phase-separated anaerobic treatment of cheese industry
wastewater. Biotechnol Bioprocess Eng 2(1):53
Imeni SM, Puy N, Ovejero J, Busquets AM, Bartroli J, Pelaz L, Ponsá S, Colón J (2020) Techno-
economic assessment of anaerobic co-digestion of cattle manure and wheat straw (raw and
pre-treated) at small to medium dairy cattle farms. Waste Biomass Valorization
11(8):4035–4051
Ince O (1998) Potential energy production from anaerobic digestion of dairy wastewater. J Environ
Sci Health Part A 33(6):1219–1228
Janczukowicz W, Zieliński M, Dębowski M (2008) Biodegradability evaluation of dairy effluents
originated in selected sections of dairy production. Bioresour Technol 99(10):4199–4205
Jürgensen L, Ehimen EA, Born J, Holm-Nielsen JB (2018) A combination anaerobic digestion
scheme for biogas production from dairy effluent—CSTR and ABR, and biogas upgrading.
Biomass Bioenergy 111:241–247
Karadag D, Koroglu OE, Ozkaya B, Cakmakci M, Heaven S, Banks C (2014) A review on
fermentative hydrogen production from dairy industry wastewater. J Chem Technol Biotechnol
89(11):1627–1636
Karthiyayini S, Sivabharathy A, Sreehari C, Sreepoorani M, Vinjth V (2017) Production of biogas
from dairy waste water. Int J Eng Res Mod Educ Special Issue 131–134
Kasmi M (2018) Biological processes as promoting way for both treatment and valorization of dairy
industry effluents. Waste Biomass Valoriz 9(2):195–209
Kolev Slavov A (2017) General characteristics and treatment possibilities of dairy wastewater—a
review. Food Technol Biotechnol 55(1):14–28
Koshta V (2010) Study on performance evaluation of anaerobic fixed film reactor for treatment of
dairy effluent using commercial packing media. Doctoral dissertation, Anand Agricultural
University, Anand
Lebrato J, Rodríguez JP, Maqueda C, Morillo E (1990) Cheese factory wastewater treatment by
anaerobic semicontinuous digestion. Resour Conserv Recycl 3(4):193–199
Li Y, Khanal SK (2016) Bioenergy: principles and applications. Wiley, Hoboken, NJ
Mainardis M, Flaibani S, Trigatti M, Goi D (2019) Techno-economic feasibility of anaerobic
digestion of cheese whey in small Italian dairies and effect of ultrasound pre-treatment on
methane yield. J Environ Manag 246:557–563
Malaspina F, Stante L, Cellamare C, Tilche A (1995) Cheese whey and cheese factory wastewater
treatment with a biological anaerobic-aerobic process. Water Sci Technol 32(12):59–72
Mawson AJ (1994) Bioconversions for whey utilization and waste abatement. Bioresour Technol
47(3):195–203
Meegoda JN, Li B, Patel K, Wang LB (2018) A review of the processes, parameters, and
optimization of anaerobic digestion. Int J Environ Res Public Health 15(10):2224
Mehrotra R, Trivedi A, Mazumdar SK (2016) Study on characterisation of Indian dairy wastewater.
Int J Eng Appl Sci Technol 1(11):77–88
Milani F, Nutter D, Thoma G (2011) Invited review: environmental impacts of dairy processing and
products: a review. J Dairy Sci 94(9):4243–4254
Mohanrao G, Subrahmanyam P (1972) Sources, flows, and characteristics of dairy wastes. Indian J
Environ Hlth 14:207–217
Motarjemi Y, Moy G, Jooste P, Anelich L (2014) Milk and dairy products. In: Food safety
management. Elsevier, pp 83–117
Murari CS, da Silva DCMN, Schuina GL, Mosinahti EF, Del Bianchi VL (2019) Bioethanol
production from dairy industrial coproducts. Bioenergy Res 12(1):112–122
6 Anaerobic Digestion of Dairy Industry Wastewater 163
Abstract Huge quantities of agricultural residues and stubbles are mainly disposed
by burning on site causing air pollution. The organic matter present in the residues
and stubble can be utilized in a planned manner, subsequently reducing the emission
(greenhouse gases) caused by burning. These agricultural stubbles are an attractive
feedstock for clean energy production through anaerobic digestion (AD). Conven-
tional liquid anaerobic digestion systems may be profitable but have a high-water
footprint. Solid-state anaerobic digestion (SSAD) not only helps to reduce water
consumption, but it also allows for a high organic loading rate and prevent nutrient
loss in the digestate. Nevertheless, process stability of an anaerobic digestion system
running on high solid concentrations may have several constraints such as limited
mass transfer and process inhibitors like ammonia, p-cresol and D-limonene if
present in the feedstock for SSAD. In the case of lignocellulosic biomass, its
recalcitrant nature may hinder the methane production under the SSAD. Apart
from these, the high total solid (TS) content may inhibit the process stability by
producing excess total volatile fatty acids (TVFAs) during SSAD.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 165
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_7
166 K. Paritosh et al.
7.1 Introduction
The modern world’s economic status and growth rate are based on energy and its
consumption (Cantarero 2020; Paritosh et al. 2020a). Depleting fossil fuel reserves,
geo-political issues over crude oil reserves, greenhouse gas (GHG) emissions and its
socio-environmental impacts are detrimental factors for a sustainable world. In the
last decades, global energy demand has increased manifold and mankind has been
forced to explore other alternative forms of energy from sustainable sources.
Renewable sources like wind, solar, ocean, small hydro, geothermal and biomass
have been accepted as major players for ensuring energy supply under a sustainable
development goal (MNRE 2019). Keeping this in mind, many countries including
the developing ones are investing in renewable energy sources. For example, total
installed grid connected renewable energy capacity in India has increased to
85.9 GW at the end of 2019 as per Ministry of New and Renewable Energy,
Government of India (MNRE 2019).
Bioenergy derived from biomass, i.e. biofuel, can be classified as liquid or
gaseous biofuels. Liquid biofuels encompass bioethanol, biodiesel and biobutanol
while example of gaseous biofuels is biomethane, biohydrogen and syngas. The
biofuels are also classified as first, second, third and fourth generation biofuels based
on the substrate used for their production. In first generation biofuels, food crops and
grains are used for biofuel production while for second generation, crop residues
such as rice straw, wheat straw, corn stover and millet straw are used. The third
generation biofuels are derived from algae and fourth generation biofuels are those
obtained from genetically modified microorganism. First generation biofuel is less
desirables as it competes with food whereas the other three biofuels are attractive to
the investors and stakeholder as they utilise renewable and waste biomass.
India produces around 634 million tonnes of agricultural stubbles on yearly basis
(Kumar et al. 2018). Organic carbon present in the agricultural stubbles and residues
may be processed for fuels and energy production. Due to the lack of effective and
efficient technology, farmers are compelled to burn these stubbles on site as to clean
it before the next crop season. This direct burning of the agricultural stubble may
produce around 1600 kg of CO2, 112 kg of CO, 9.2 kg of CH4, 5 kg of particulate
matter and 6 kg of hydrocarbons per ha land (Guo et al. 2020). Theoretically,
burning of these stubbles not only contributes to high GHG emissions, but also
causes immediate problems in the surrounding areas in the form of severe deterio-
ration of air quality (smog formation), and hence crop burning is not at all a
sustainable approach for its disposal.
Agricultural stubbles have huge energy potential and may substitute fossils for
fuel or electricity and are a promising alternative to meet future energy demands
(Hansen et al. 2020). Presently, bio-based energy has approximately 15% share in
the Worlds’ total energy use which is almost 45 EJ. Numerous studies have
suggested that the potential market for bioenergy may increase up to 50% of the
total energy use by the year 2050 (Perea-Moreno et al. 2019). As per a study,
220 billion tonnes of dry biomass are produced worldwide annually (Dahunsi and
7 Solid State Anaerobic Digestion of Agricultural Waste for Bioenergy Production 167
and improve syntrophic relations in the reactor. Moreover, the presence of materials
such as biochar or activated carbon accelerate the utilization of volatile fatty acids
(VFA) and ensure availability of substrate to methanogens. Whereas, addition of
elements such as cobalt (Co), nickel (Ni), iron (Fe) and zinc (Zn). in the AD acts as
micronutrients and accelerate the metabolic activities of methanogens which pro-
vides a better yield of methane (Paritosh et al. 2020b). Nanoparticles of the above
mentioned trace elements have also been supplied by various researches to enhance
anaerobic digestion of biomass and biogas production (Lee and Lee 2019).
Anaerobic digestion can be categorized into two distinct forms based on their total
solid (TS) content in the reactor medium. The first one is liquid state anaerobic
digestion with a solid content <15%, whereas the other one being solid state
anaerobic digestion (SSAD) with a TS content >15%. SSAD has the following
advantages over liquid anaerobic digestion (LAD): feasibility of using higher
organic loading rate (OLR), less energy requirement, smaller reactor volume and
increased volumetric methane yield (Brown et al. 2012; Rico et al. 2015; Panjičko
et al. 2017). Beside, pathogen inactivation may also be achieved in SSAD of
biodegradable waste (Jiang et al. 2018).
However, SSAD has a few challenges which include slow mass transfer, process
instability, end product needs additional treatment and lower biogas production
(Karthikeyan and Visvanathan 2013; Carlos-Pinedo et al. 2019). These issues need
to be addressed in order to enhance process efficiency, and to further ensure its
feasibility at a larger scale for successful commercialization of this technology.
Several types of feedstocks including various wastes such as the organic fraction of
municipal solid waste (OFMSW), food waste, forest waste, agricultural waste,
animal waste, solid manure, energy crops, industrial waste, residual lignocellulosic
biomass, paper and pulp waste have been established as good substrates for biogas
generation using the SSAD process (Fig. 7.1). The physical and chemical
SSAD
Feedstock
Agro-
Crop
Food waste Energy crops OFMSW industry
residues
waste
composition of the substrate are very important and directly affect the process
efficiency of AD system. For example, the presence of a high amount of recalcitrant
compounds (such as lignin) in the biomass lower the biogas production whereas a
high amount of easily utilizable compounds (such as sugars) enhances methanogenic
activity (Paritosh et al. 2019). Hence, even before starting the process the suitability
of any substrate, including lignocellulosic biomass for SSAD, should be determined
by biomass characterization.
The organic fraction of municipal solid waste (OFMSW) comprises of yard trim-
mings, grass clippings, vegetable wastes, food wastes and fruit peel wastes
(Kesharwani and Bajpai 2020). The approach for waste collection and transportation
plays a major role in the SSAD process performance. Also, the seasonal variations
and environment conditions can affect the composition of the OFMSW which in turn
affects the SSAD process performance (Forster-Carneiro et al. 2007). Several studies
have demonstrated that both physical as well as chemical properties of OFMSW
have significant impact on biogas production.
Michele et al. (2015) performed SSAD of OFMSW by recirculation of the
digestate. This liquid digestate flushing helped in removing fermentative products
(such as volatile fatty acids (VFA)) inhibiting methanogenesis. The ratio of solid
waste to digestate was in between 1:1.18–1:0.9 on w/w basis. The total solids
removal was 36.9%, however the loss of organic matter was attributed to the
washout with the percolate from the reactor. Hence, the percolate which was high
in organic content was subjected to AD for biogas production in a second AD reactor
(LAD). The mass balance showed that the methane content from the dry AD and the
percolate were 18.4% and 49.7%, respectively, at a 21 d hydraulic retention time
(HRT). However, only 20.4% and 25.7% of potential producible methane was
generated by adopting 15 and 20 d of HRT using LAD of the same waste.
Food waste is also considered a part of OFMSW and contains organic materials
which are transformed into simple molecules that are readily digested in the AD
process. However, accumulation of VFAs caused by high soluble organic contents
act as inhibitor by decreasing the pH of the system leading to reduction in methane
yield of the AD process (Micolucci et al. 2018). Co-digestion of OFMSW with
lignocellulosic biomass can be a beneficial approach for enhancing the process
efficiency. Brown and Li (2013) examined the effect of feedstock to inoculum
(F/I) ratio (1, 2 and 3) and substrate concentration (0, 10 and 20%) on co-digestion
of food waste (FW) with yard waste on biogas production using SSAD. A high
volumetric biogas production rate (8.6 L per L reactor volume) was achieved with
10% FW concentration and a F/I value of 2.
In another study conducted by Wang et al. (2012), the effect of different ratios of
FW to distiller’s grain on biogas production using SSAD was investigated. A 75.7%
increase in the biogas production was observed with co-digestion compared with
170 K. Paritosh et al.
study, Liew et al. (2012) compared the biogas production potential of corn stover
with yard waste, leaves and wheat straw using SSAD at F/I ratio of 2. The maximum
methane yield of 81.2 L/kg VS was for corn stover as feedstock, while yard waste,
leaves and wheat straw yielded 40.8, 55.4 and 66.9 L/kg VS of methane,
respectively.
Methane production from albizia plant biomass was examined in two different
anaerobic digestion systems, i.e. SSAD and LAD rectors (Ge et al. 2014). The study
found higher methane production using LAD where the methane yield from albizia
leaves and wood chips were 161 and 113 L/kg VS, respectively. The methane
production from albizia leaves using SSAD was comparable (156.8 L/kg VS) to
the LAD system, however, it was much lower (59.6 L/kg VS) in case of albizia wood
chips using SSAD. Similar to other previously mentioned studies, the volumetric
methane production was much higher (five times more) in the SSAD system in
comparison to the LAD reactor.
Cui et al. (2011) compared raw wheat straw and spent wheat straw from horse
stall for biogas production in a SSAD system. The experimental conditions used
during the study were 20% TS, F/I ratio of 2, 4 and 6, and the inoculum used was
digestate collected from a LAD reactor. The maximum daily methane yield from
spent wheat straw was observed 8 and 3 days earlier in comparison to raw wheat
straw with a F/I ratio of 2 and 4, respectively, indicating improved degradation rate
for spent wheat straw. The maximum methane generation of 150 L/kg was with
spent wheat straw when the F/I ratio was 4 and it was 56.2% higher than that of raw
straw. Cellulose and hemicellulose digestibility was also, respectively, 24.1 and
49.4% higher in spent wheat straw compared with raw straw.
Yan et al. (2015) investigated the effects of different parameters such as solid
concentration, temperature and C/N ratio on the digestion of rice straw employed for
biogas production using a SSAD system. Maximum biogas production (447.4 mL/g
VS) was observed with an initial TS of 20% and C/N ratio of 29.6 at 35.6 oC.
Sheets et al. (2015) investigated the influence of different factors, namely TS
concentration (20 and 30%), temperature (36 and 55 C) and controlled air exposure,
on biogas production using switch grass as the substrate in a SSAD system. The air
exposure did not show any positive effect on the methane production from switch
grass. The biogas generation increased from 102 and 145 L CH4/kg of VS with
increase in TS concentration from 20 and 30% in mesophilic conditions. Under
thermophilic conditions, the methane yields were 88 and 113 L CH4/kg VS for
20 and 30% of solid concentrations, respectively.
Contrary to this previous study, biogas production decreased with increase in
initial TS concentration from 8 to 38% during biogas production from giant reed
biomass (Yang and Li 2014). The inhibitory effect was attributed to the high
accumulation of VFA at high solids concentration. The maximum methane produc-
tion (129.7 L CH4/kg VS) was at a F/I ratio of 2 and TS content of 20–23% using the
SSAD process (Table 7.1).
7 Solid State Anaerobic Digestion of Agricultural Waste for Bioenergy Production 173
Table 7.1 Methane production from lignocellulosic waste using SSAD process
TS T CH4
Feedstock (%) (oC) yield Remarks Reference
Rice straw 20 37 263 Incubation time and moisture significantly Mustafa
L/kg affected the lignin degradation et al.
VS (2016)
Wheat 18 37 254 Fungal treatment facilitated faster start-up of Rouches
straw L/kg SS–AD reactor et al.
VS (2019)
Rice straw 20 37 258 Fungal treatment showed linear relation Mustafa
L/kg between methane yield and lignin et al.
VS degradation (2017)
Palm fruit 20 40 73.3 Straw mushroom cultivation reduced the Mamimin
bunches m3 / recalcitrance et al.
tonne (2021)
Sugarcane 15 35 143 Lignin droplets formed during thermal Lima et al.
bagasse L/kg treatment hindered the hydrolysis (2018)
VS
Rice husk 21 – 18 Optimization of enzyme concentration is Nugraha
L/kg required et al.
TS (2018)
Distilled 20 52 212 Methanoculleus and Methanosarcina were Wang
grain L/kg detected in abundance et al.
VS (2018)
Rice straw 21 37 190 68% higher glucose yield was obtained at Momayez
L/kg 60 min treatment et al.
VS (2018)
Wheat 25 35 1.2 Startup time of SS–AD digester was reduced Zhu et al.
straw m3 / by 10 days (2020)
m3d
Rice straw 15 35 357 Gas productivity was improved by 2.85– Qian et al.
+ manure L/kg 5.88% per unit TS after treatment (2019)
TS
Rose stalk 12.1 55 117 Treatment facilitated higher VS removal and Liang et al.
L/kg lower digestion time (2016)
VS
Solid concentration is one of the most important parameters for the SSAD process
and significantly affects the process efficiency. Hence, many of the previous studies
have focused on optimizing the solids concentration in the digester. A very high
solid concentration in the SSAD process contributes to reduced biogas production by
limiting microbial access to the substrate (Bollon et al. 2013). The water content in
the system is also relevant in this regard as it facilitates mass transfer and low water
174 K. Paritosh et al.
content can suppress the digestion process in the SSAD system (Le Hyaric et al.
2012).
Anaerobic digestion of municipal solid waste (MSW) was studied at two different
solid concentrations of 20% and 30% under mesophilic conditions (Fernández et al.
2010). The dissolved organic carbon and VFAs removal was higher at low TS
concentration of 20%, whereas at high TS concentration (30%) digestion of organic
waste compounds decreased by 17%. Abbassi-Guendouz et al. (2012) investigated
digestion of cardboard at various solid concentrations (10–30%). The results dem-
onstrated that increase in the solid concentration was detrimental to the methane
production rate. The threshold value for TS was 30% in this study, and beyond this
methanogenic activity gets inhibited. In another similar study on methane production
from organic wastes obtained from the palm oil industry (oil farm fronds, oil palm
trunks and empty fruit branches) at three different solid contents (16, 25 and 35%)
observed a negative correlation with increase in solid concentration in the AD
process (Suksong et al. 2016). The maximum methane production (72 L/kg biomass)
and total solids removal was at 16% solid concentration, whereas the methane yield
decreased for the other two solid contents.
Hence, from the above studies it can be summarized that methane yield and
methanogenic activity tends to decrease with an increase in solid concentration. The
reason behind this trend is mainly related to the dysfunction of mass transfer at high
solid content (Abbassi-Guendouz et al. 2012; Fernández et al. 2010). For example,
Bollon et al. (2013) found that when solid concentration increased from 10% to 25%,
the medium solutes diffusion coefficient reduced by 3.7 times.
7.5.2 Inoculum
Inoculum is another important factor as it provides the microbes, the main catalyst in
the AD process (Cui et al. 2011; Shi et al. 2014). LAD effluents and digestate from
the SSAD process are generally better inocula than activated sludge, rumen fluid and
manure because the digestate from anaerobic processes provides high numbers of
active methanogens that are more suited to the AD process. For example, Xu et al.
(2016) established in their study that effluent from the LAD process is a better
inoculum source than manure, rumen fluid, lake sediments and sewage sludge for
initiating the SSAD process. In another study, Forster-Carneiro et al. (2007) noted
that the lag phase in the SSAD process reduced from 20–30 days to 2–5 days when
LAD effluent was used as inoculum instead of fresh manure. Suksong et al. (2019)
reported a twofold increase in methane yield using LAD effluent as inoculum in
comparison to SSAD finished materials. The LAD effluent used had high alkalinity
(5.9 g/kg) and low VFA concentration (0.05 g/kg) which may have contributed to
the better performance of the system (Suksong et al. 2019).
Often recalcitrant components in biomass prevent efficient utilization of the
biomass for biogas production. In such cases, different process improvement strat-
egies are applied, one among them is the use of hydrolytic microorganisms. Weiß
7 Solid State Anaerobic Digestion of Agricultural Waste for Bioenergy Production 175
et al. (2010) used enriched hydrolytic microbes for enhanced degradation of ligno-
cellulosic biomass rich in hemicellulose. The study found an increase in xylanase
activity by 1.62% as well as 53% increase in methane yield with supplementation of
hemicellulolytic bacteria to the AD process. According to Ma et al. (2013) the
optimal ratio of hydrolytic microbes to methanogens was recommended to be
24 in AD process, the hydrolysis process becomes the rate limiting step at a ratio
below 24, while a ratio higher than 24 makes methanogenesis the rate limiting step.
Similarly, enhancement in biogas production from corn stover due to the addition of
dairy manure as inocula was attribute to the activity of hydrolytic microbes in the
AD process (Xu et al. 2013). The biogas yield from corn stover using dairy manure
was 30% and 100% higher than those using sewage sludge finished material and
food waste as inoculum.
Gu et al. (2014) compared different inoculum sources such as digestates from
dairy manure, chicken manure, municipal sludge, swine manure, paper mill sludge
and anaerobic granular sludge for biogas production with rice straw as the substrate.
Compared to sludge, digested manure as inoculum demonstrated significantly
improved lignocellulose degradation and methane production due to the high
enzyme activity (mainly cellulase and xylanase) in animal manure digestates.
The inoculation size in SSAD is another aspect which has the ability to increase
methanogenic activity. The optimized concentration of inoculum can give a good
start to the SSAD process and may as well reduce the lag phase of the AD process
significantly (Yang et al. 2015). The inoculum size in AD is often described as food
to inoculum (F/I) ratio. At mesophilic conditions, inoculation size as F/I ratio of 2–3
on VS basis is recommended for the AD process of lignocellulosic biomass (Zhu
et al. 2014; Liew et al. 2012; Ma et al. 2013). Under thermophilic range, the optimal
F/I ratio should be in the range of 4–6 when the experiment was performed on corn
stover. This difference in optimum F/I ratio under different temperature conditions
was also confirmed by Li et al., where the maximum methane yield for mesophilic
and thermophilic conditions was at F/I ratio of 2.43 and 4.58, respectively. Lin et al.
(2015) investigated SSAD of yard trimmings comprised of wood chips, maple leaves
and lawn grass as substrate for biogas production and found a F/I ratio of 4–6 to be
better for the digestion process under thermophilic conditions (55 oC). In another
study, the F/I ratio of 1 showed best results for methane production under mesophilic
temperature (Brown and Li 2013).
Mixing of inoculum with the substrate is another important aspect of the SSAD
process. In this regard, mixing of inoculum with the substrate is required prior to the
loading in the SSAD reactors. This pre-mixing is particularly needed in case of
processes with high solid content. In large or pilot scale SSAD bioreactors, the
interaction between microbes and feedstock sometimes fails due to improper mixing.
Two different scenarios were created by Zhu et al. (2014) for analysing the effect of
premixing and partial mixing on SSAD process stability and net methane yield. In
the first scenario, the whole inoculum was completely mixed with the substrate at the
start of the process. In the second scenario, half of the inoculum was mixed with
substrate, following which the rest of the 50% inoculum was poured onto the top.
Although, the methane yield was the same in both scenarios, the start-up time was
176 K. Paritosh et al.
less in the premixed SSAD reactor. In another study, three premixing strategies were
employed to digest corn stover anaerobically in a SSAD reactor (Zhu et al. 2014).
Comparison of the completely mixed scenario with partially mixed in one layer and
two layers was performed. The reactor with two layered partial mixing of inoculum
yielded the highest methane at F/I ratio of 4 to 6.
7.5.3 Temperature
Temperature is one of the most important determining factors for the growth and
survival of microbes in the AD process at both laboratory and industrial scale
systems (de Diego-Díaz et al. 2018). Reactor temperature can selectively enrich
microbes and has the capacity to enhance the rate of biochemical reactions in the
bioreactor. The temperature ranges used for the AD process are as follows: thermo-
philic (55–70 oC), mesophilic (20–45 oC) and psychrophilic (0–20 C). Among
these temperature ranges thermophilic and mesophilic conditions have been exten-
sively practiced for the degradation process of lignocellulosic biomass (LCB) and
OFMSW in SSAD. The mesophilic temperature range is more preferred when
compared to thermophilic temperatures due to greater process stability as well as
better growth of methanogens. Although the thermophilic temperature zone has its
own benefits in the AD process, it requires more energy input in the process, making
the process economics unsustainable. However, Sheets et al. (2015) during SSAD of
switch grass concluded that under thermophilic conditions, net energy input can be
decreased with the increase in methane production rate.
Furthermore, thermophilic temperature accelerates the process at initial level and
drives the hydrolysis faster, but often methanogenic conversion is not satisfactory
(Yang et al. 2015). Hydrolysis of substrate can be accelerated in thermophilic
conditions due to enrichment of hydrolytic microorganisms inside the SSAD biore-
actor. But faster hydrolysis of biomass often results in volatile fatty acids (VFAs)
accumulation in the system, causing acidification of the reactor (Shi et al. 2014).
This acidification further reduces methanogenesis, decreasing biogas production and
also reducing stability of the SSAD system (Yan et al. 2015).
Shi et al. (2014) reported that the degradation rate of cellulose and hemicellulose
was higher under thermophilic conditions in contrast to the mesophilic temperature
range. In another study, a total 6–41% of cellulose and 2–34% of hemicelluloses
digestion was observed during thermophilic SSAD of lignocellulosic biomass.
These improved results were attributed to the increased (10–50 times) presence of
cellulolytic and xylanolytic microorganisms in the thermophilic SSAD bioreactor
(Fernández-Rodríguez et al. 2013).
7 Solid State Anaerobic Digestion of Agricultural Waste for Bioenergy Production 177
7.5.4 Inhibition
There are many factors that can cause inhibition in the methanogenesis process in
SSAD. For example, excess VFA accumulation can greatly affect methanogens,
causing instability in the bioreactor (Carlos-Pinedo et al. 2019). Acidification results
in decreased pH values, thus inhibiting methanogens which are most susceptible to
the environmental conditions (Rocamora et al. 2020). The significant reason behind
the increment in VFA accumulation in anaerobic digestion reactors is feedstock
overloading (Eko and Chaiprasert 2020). Zhang et al. demonstrated that the use of
alternative feedstock can avoid VFA accumulation for better stability of the AD
process. The addition of packaging waste along with FW can avoid VFA accumu-
lation during the SSAD process. The study suggested that choice of heterogenous
waste as feedstock may permit high loading of substrate during the digestion. The
ratio of VFA to alkalinity can assist to regulate digester stability. A VFA/alkalinity
ratio within 0.3–0.4 is generally observed in AD plants, but a ratio in the range
0.4–0.6 can provide a stable and safe operation when high organic containing
substrates are used (Lossie and Pütz 2008).
Besides VFA accumulation and alkalinity, the ammonia nitrogen content can also
bring instability in the AD process. A study conducted by Duan et al. (2012) on
sewage sludge found reduced methane generation even at a VFA/alkalinity ratio of
0.2 due to excessive ammonia nitrogen concentrations. This demonstrates that
measuring the VFA/alkalinity ratio to monitor reactor condition could be deceptive
in the long term operation of SSAD. A suitable knowledge of ammonium inhibition
is required to predict the process steadiness.
Free ammonia (NH3) and ammonium ion (NH4+) are available during the diges-
tion of nitrogenous matter and feedstocks rich in protein (FW and OFMSW). The
concentration of the ionic form as well as the non-ionic form of ammonia is
influenced by both temperature and pH of the SSAD system as described by the
following equations (7.1) and (7.2) (Calli et al. 2005).
2729:92
pK a ¼ 0:09018 ð7:1Þ
T þ 273:15
TAN
FAN ¼ ð7:2Þ
1 þ 10ðpK a pH Þ
Lignin present in lignocellulosic biomass is inhibitory to the SSAD process due to its
recalcitrant nature. In order to increase the production of biogas and reduce inhibi-
tion, different pretreatment methods can be applied (Kumar et al. 2018; Saha et al.
2018). Chemical pretreatment involves acid, alkali, ionic liquids (ILs) and organic
solvents to disrupt linkage between complexes in the lignocellulosic matrix (Kumar
et al. 2018). Whereas, physiochemical pretreatment involves usage of carbon diox-
ide explosion, ammonia fibre explosion (AFEX) and wet oxidation. AFEX treatment
includes pressurized ammonia given to biomass with rapid decompression (Stoklosa
et al. 2017). As a result, hydrolysis and ammonolysis reactions break the ester cross
links in the cell wall biopolymers. With the help of biomass pretreatment, various
advantages can be achieved such as lignin removal, decrystallization of cellulose,
increase accessible surface area, alteration of inter-linkage of hemicelluloses and
cellulose in biomass structure (Rouches et al. 2019). The cellulose decrystallization
causes cellulose to become more porous and readily available to the microbes, which
enhances its bioconversion efficiency (Paritosh et al. 2021; Yadav et al. 2019).
Pretreatment for decrystallization of cellulose before digestion can be carried out
with the help of acids. Inorganic acids such as hydrochloric acid (HCl), sulphuric
acids (H2SO4) and phosphoric acid (H3PO4) are commonly employed for this
purpose. However, in the recent times ionic liquids (ILs) have also been used for
7 Solid State Anaerobic Digestion of Agricultural Waste for Bioenergy Production 179
biomass pretreatment. ILs are less corrosive in nature, connect with the hydroxyl
group of cellulose by breaking hydrogen bonds and this ensures dissolution of
cellulose (Han et al. 2020). The pretreatment process using ionic liquids is efficient
in recovering decrystallized cellulose with the help of anti-solvents such as metha-
nol, acetone, ethanol or water and also, the ILs can be recovered to a very high extent
(even 100% in some cases) (Han et al. 2020).
The most preferred ionic liquid used for pretreatment of lignin containing bio-
mass is N-methyl morpholine-N-oxide monohydrate (NMMO). Akhand and
Méndez Blancas (2012) reported a total of 47% increase in methane yield when
rice straw biomass was subjected to NMMO pretreatment. The pretreatment
increased the substrate surface area which facilitated increased microbial degrada-
tion of the feedstock to produce biogas.
Physical pretreatment such as size reduction was applied for methane production
from napier grass with three sizes of 6, 10 and 20 mm (biomass passing through
respective size sieves) (Surendra and Khanal 2015). A higher methane yield was
found for the smallest biomass size of 6 mm as compared to the two other biomass
sizes (10 and 20 mm). This improved results is again attributed to the increase in
specific surface area for microbial degradation of biomass.
Various pretreatment methods such as steam explosion, irradiation, dilute acid
application and liquid hot water have been developed to enhance biogas production
and reduce inhibition (Kumar et al. 2018). In addition, other methods such as wet
oxidation, alkaline treatment and biological methods (fungal or enzymatic) can be
applied for lignin removal (Kumar et al. 2018). Zhao et al. (2014) investigated
pretreatment of yard trimmings using white rot fungi (Ceriporiopsis subvermispora)
for improving the SSAD process. Ceriporiopsis subvermispora pretreatment at 40%
solid concentration showed the highest methane production (44.6 L/kg VS) which
was 154% higher than methane produced from raw yard trimmings. Similarly, when
albizia chips were pretreated with the same fungal strain of Ceriporiopsis
subvermispora, 370% increase in biogas yield was reported (Ge et al. 2015).
Pretreatment of rice straw with combined physical (milling) and biological (fungal)
methods for improved biodegradability of feedstock in the SSAD system was
studied (Mustafa et al. 2017). A 1 month long incubation with Pleurotus ostreatus
and subsequent milling of the rice straw achieved 30.4% lignin removal and 165%
higher methane production in comparison to the experiments with untreated rice
straw.
However, to degrade a higher lignin content in feedstocks such as spruce (29%
lignin content), the alkaline pretreatment method is more suited. In a study by
Mohsenzadeh et al. (2012), birch and spruce biomass was pretreated with different
alkaline reagent combinations (NaOH/urea, NaOH/thiourea, NaOH/urea/thiourea,
and NaOH/polyethylene glycol) at four different temperatures (15, 0, 22 and
80 oC). The pretreament with combinations of NaOH/thiourea at 15 oC showed
the best results in terms of 59.9% and 45.3% increase in yield using birch and spruce
biomass, respectively. Although lignin removal was not maximum at this
pretreatment condition, product yield was the highest, indicating other factors such
as crystallinity of sugars in the biomass have more significance. According to the
180 K. Paritosh et al.
References
Eko H, Chaiprasert P (2020) Enhancement of methane production from high solid anaerobic
digestion of pretreated palm oil decanter cake using a modified solid inclined reactor. J Chem
Technol Biotechnol 95(3):781–790
Fernández J, Pérez M, Romero LI (2010) Kinetics of mesophilic anaerobic digestion of the organic
fraction of municipal solid waste: influence of initial total solid concentration. Bioresour
Technol 101(16):6322–6328
Fernández-Rodríguez J, Pérez M, Romero LI (2013) Comparison of mesophilic and thermophilic
dry anaerobic digestion of OFMSW: kinetic analysis. Chem Eng J 232:59–64
Forster-Carneiro T, Pérez M, Romero LI, Sales D (2007) Dry-thermophilic anaerobic digestion of
organic fraction of the municipal solid waste: focusing on the inoculum sources. Bioresour
Technol 98(17):3195–3203
Ge X, Matsumoto T, Keith L, Li Y (2014) Biogas energy production from tropical biomass wastes
by anaerobic digestion. Bioresour Technol 169:38–44
Ge X, Matsumoto T, Keith L, Li Y (2015) Fungal pretreatment of Albizia chips for enhanced biogas
production by solid-state anaerobic digestion. Energy Fuel 29(1):200–204
Gu Y, Chen X, Liu Z, Zhou X, Zhang Y (2014) Effect of inoculum sources on the anaerobic
digestion of rice straw. Bioresour Technol 158:149–155
Guo L, Ma Y, Tigabu M, Guo X, Zheng W, Guo F (2020) Emission of atmospheric pollutants
during forest fire in boreal region of China. Environ Pollut 264:114709
Han SY, Park CW, Kwon GJ, Kim NH, Kim JC, Lee SH (2020) Ionic liquid pretreatment of
lignocellulosic biomass. J For Environ Sci 36(2):69–77
Hansen JH, Hamelin L, Taghizadeh-Toosi A, Olesen JE, Wenzel H (2020) Agricultural residues
bioenergy potential that sustain soil carbon depends on energy conversion pathways. GCB
Bioenergy 12(11):1002–1013. https://mnre.gov.in/img/documents/uploads/file_f-
1597797108502.pdf
Jain S, Jain S, Wolf IT, Lee J, Tong YW (2015) A comprehensive review on operating parameters
and different pretreatment methodologies for anaerobic digestion of municipal solid waste.
Renew Sustain Energy Rev 52:142–154
Jiang Y, Dennehy C, Lawlor PG, Hu Z, Zhan X, Gardiner GE (2018) Inactivation of enteric
indicator bacteria and system stability during dry co-digestion of food waste and pig manure. Sci
Total Environ 612:293–302
Karthikeyan OP, Visvanathan C (2013) Bio-energy recovery from high-solid organic substrates by
dry anaerobic bio-conversion processes: a review. Rev Environ Sci Biotechnol 12(3):257–284
Kesharwani N, Bajpai S (2020) Batch anaerobic co-digestion of food waste and sludge: a multi
criteria decision modelling (MCDM) approach. SN Appl Sci 2(8):1–11
Kumar S, Paritosh K, Pareek N, Chawade A, Vivekanand V (2018) De-construction of major Indian
cereal crop residues through chemical pretreatment for improved biogas production: an over-
view. Renew Sustain Energy Rev 90:160–170
Le Hyaric R, Benbelkacem H, Bollon J, Bayard R, Escudié R, Buffière P (2012) Influence of
moisture content on the specific methanogenic activity of dry mesophilic municipal solid waste
digestate. J Chem Technol Biotechnol 87(7):1032–1035
Lee YJ, Lee DJ (2019) Impact of adding metal nanoparticles on anaerobic digestion performance–a
review. Bioresour Technol 292:121926
Liang YG, Cheng B, Si YB, Cao DJ, Li DL, Chen JF (2016) Effect of solid-state NaOH
pretreatment on methane production from thermophilic semi-dry anaerobic digestion of rose
stalk. Water Sci Technol 73(12):2913–2920
Liew LN, Shi J, Li Y (2012) Methane production from solid-state anaerobic digestion of lignocel-
lulosic biomass. Biomass Bioenergy 46:125–132
Lima DRS, Adarme OFH, Baêta BEL, Gurgel LVA, de Aquino SF (2018) Influence of different
thermal pretreatments and inoculum selection on the biomethanation of sugarcane bagasse by
solid-state anaerobic digestion: a kinetic analysis. Ind Crop Prod 111:684–693
Lin Y, Ge X, Liu Z, Li Y (2015) Integration of Shiitake cultivation and solid-state anaerobic
digestion for utilization of woody biomass. Bioresour Technol 182:128–135
182 K. Paritosh et al.
Lossie U, Pütz P (2008) Targeted control of biogas plants with the help of FOS/TAC. Practice
Report Hach-Lange
Ma J, Frear C, Wang ZW, Yu L, Zhao Q, Li X, Chen S (2013) A simple methodology for rate-
limiting step determination for anaerobic digestion of complex substrates and effect of microbial
community ratio. Bioresour Technol 134:391–395
Maletta E, Díaz-Ambrona CH (2020) Lignocellulosic crops as sustainable raw materials for
bioenergy. Green Energy Sustain Strateg Global Ind:489–514
Mamimin C, Chanthong S, Leamdum C, Sompong O, Prasertsan P (2021) Improvement of empty
palm fruit bunches biodegradability and biogas production by integrating the straw mushroom
cultivation as a pretreatment in the solid-state anaerobic digestion. Bioresour Technol
319:124227
Michele P, Carlo M, Sergio S, Fabrizio A (2015) Optimization of solid state anaerobic digestion of
the OFMSW by digestate recirculation: a new approach. Waste Manag 35:111–118
Micolucci F, Gottardo M, Pavan P, Cavinato C, Bolzonella D (2018) Pilot scale comparison of
single and double-stage thermophilic anaerobic digestion of food waste. J Clean Prod
171:1376–1385
MNRE (2019) Annual report 2019–20. Ministry of New and Renewable Energy. Government of
India. www.mnre.gov.in. Accessed 10 Dec 2020
Mohsenzadeh A, Jeihanipour A, Karimi K, Taherzadeh MJ (2012) Alkali pretreatment of softwood
spruce and hardwood birch by NaOH/thiourea, NaOH/urea, NaOH/urea/thiourea, and NaOH/
PEG to improve ethanol and biogas production. J Chem Technol Biotechnol 87(8):1209–1214
Momayez F, Karimi K, Horváth IS (2018) Enhancing ethanol and methane production from rice
straw by pretreatment with liquid waste from biogas plant. Energ Conver Manage 178:290–298
Mustafa AM, Poulsen TG, Sheng K (2016) Fungal pretreatment of rice straw with Pleurotus
ostreatus and Trichoderma reesei to enhance methane production under solid-state anaerobic
digestion. Appl Energy 180:661–671
Mustafa AM, Poulsen TG, Xia Y, Sheng K (2017) Combinations of fungal and milling pretreat-
ments for enhancing rice straw biogas production during solid-state anaerobic digestion.
Bioresour Technol 224:174–182
Nugraha WD, Keumala CF, Matin HHA (2018) The effect of acid pre-treatment using acetic acid
and nitric acid in the production of biogas from rice husk during solid state anaerobic digestion
(SS-AD). In E3S web of conferences (Vol 31), p 01006. EDP Sciences
Panjičko M, Zupančič GD, Fanedl L, Logar RM, Tišma M, Zelić B (2017) Biogas production from
brewery spent grain as a mono-substrate in a two-stage process composed of solid-state
anaerobic digestion and granular biomass reactors. J Clean Prod 166:519–529
Paritosh K, Pareek N, Chawade A, Vivekanand V (2019) Prioritization of solid concentration and
temperature for solid state anaerobic digestion of pearl millet straw employing multi-criteria
assessment tool. Sci Rep 9(1):1–11
Paritosh K, Balan V, Vijay VK, Vivekanand V (2020a) Simultaneous alkaline treatment of pearl
millet straw for enhanced solid state anaerobic digestion: experimental investigation and energy
analysis. J Clean Prod 252:119798
Paritosh K, Yadav M, Chawade A, Sahoo D, Kesharwani N, Pareek N, Vivekanand V (2020b)
Additives as a support structure for specific biochemical activity boosts in anaerobic digestion: a
review. Front Energy Res 8:88
Paritosh K, Yadav M, Kesharwani N, Pareek N, Karthyikeyan OP, Balan V, Vivekanand V (2021)
Strategies to improve solid state anaerobic bioconversion of lignocellulosic biomass an over-
view. Bioresour Technol 125036
Perea-Moreno MA, Samerón-Manzano E, Perea-Moreno AJ (2019) Biomass as renewable energy:
worldwide research trends. Sustainability 11(3):863
Qian X, Shen G, Wang Z, Zhang X, Chen X, Tang Z, Lei Z, Zhang Z (2019) Enhancement of high
solid anaerobic co-digestion of swine manure with rice straw pretreated by microwave and
alkaline. Bioresour Technol Rep 7:100208
7 Solid State Anaerobic Digestion of Agricultural Waste for Bioenergy Production 183
Rico C, Montes JA, Muñoz N, Rico JL (2015) Thermophilic anaerobic digestion of the screened
solid fraction of dairy manure in a solid-phase percolating reactor system. J Clean Prod
102:512–520
Rocamora I, Wagland ST, Villa R, Simpson EW, Fernández O, Bajón-Fernández Y (2020) Dry
anaerobic digestion of organic waste: a review of operational parameters and their impact on
process performance. Bioresour Technol 299:122681
Rouches E, Escudié R, Latrille E, Carrère H (2019) Solid-state anaerobic digestion of wheat straw:
impact of S/I ratio and pilot-scale fungal pretreatment. Waste Manag 85:464–476
Ruiz B, Flotats X (2014) Citrus essential oils and their influence on the anaerobic digestion process:
an overview. Waste Manag 34(11):2063–2079
Saha S, Jeon BH, Kurade MB, Jadhav SB, Chatterjee PK, Chang SW, Govindwar SP, Kim SJ
(2018) Optimization of dilute acetic acid pretreatment of mixed fruit waste for increased
methane production. J Clean Prod 190:411–421
Sheets JP, Ge X, Li Y (2015) Effect of limited air exposure and comparative performance between
thermophilic and mesophilic solid-state anaerobic digestion of switchgrass. Bioresour Technol
180:296–303
Shi J, Xu F, Wang Z, Stiverson JA, Yu Z, Li Y (2014) Effects of microbial and non-microbial
factors of liquid anaerobic digestion effluent as inoculum on solid-state anaerobic digestion of
corn stover. Bioresour Technol 157:188–196
Stoklosa RJ, del Pilar Orjuela A, da Costa Sousa L, Uppugundla N, Williams DL, Dale BE, Hodge
DB, Balan V (2017) Techno-economic comparison of centralized versus decentralized
biorefineries for two alkaline pretreatment processes. Bioresour Technol 226:9–17
Suksong W, Kongjan P, Prasertsan P, Imai T, Sompong O (2016) Optimization and microbial
community analysis for production of biogas from solid waste residues of palm oil mill industry
by solid-state anaerobic digestion. Bioresour Technol 214:166–174
Suksong W, Mamimin C, Prasertsan P, Kongjan P, Sompong O (2019) Effect of inoculum types
and microbial community on thermophilic and mesophilic solid-state anaerobic digestion of
empty fruit bunches for biogas production. Ind Crop Prod 133:193–202
Surendra KC, Khanal SK (2015) Effects of crop maturity and size reduction on digestibility and
methane yield of dedicated energy crop. Bioresour Technol 178:187–193
Thanh PM, Ketheesan B, Yan Z, Stuckey D (2016) Trace metal speciation and bioavailability in
anaerobic digestion: a review. Biotechnol Adv 34(2):122–136
Wang LH, Wang Q, Cai W, Sun X (2012) Influence of mixing proportion on the solid-state
anaerobic co-digestion of distiller’s grains and food waste. Biosyst Eng 112(2):130–137
Wang TT, Sun ZY, Huang YL, Tan L, Tang YQ, Kida K (2018) Biogas production from distilled
grain waste by thermophilic dry anaerobic digestion: pretreatment of feedstock and dynamics of
microbial community. Appl Biochem Biotechnol 184(2):685–702
Weiß S, Tauber M, Somitsch W, Meincke R, Müller H, Berg G, Guebitz GM (2010) Enhancement
of biogas production by addition of hemicellulolytic bacteria immobilised on activated zeolite.
Water Res 44(6):1970–1980
Xu F, Shi J, Lv W, Yu Z, Li Y (2013) Comparison of different liquid anaerobic digestion effluents
as inocula and nitrogen sources for solid-state batch anaerobic digestion of corn stover. Waste
Manag 33(1):26–32
Xu F, Wang F, Lin L, Li Y (2016) Comparison of digestate from solid anaerobic digesters and
dewatered effluent from liquid anaerobic digesters as inocula for solid state anaerobic digestion
of yard trimmings. Bioresour Technol 200:753–760
Xu F, Li Y, Ge X, Yang L, Li Y (2018) Anaerobic digestion of food waste—challenges and
opportunities. Bioresour Technol 247:1047–1058
Yabu H, Sakai C, Fujiwara T, Nishio N, Nakashimada Y (2011) Thermophilic two-stage dry
anaerobic digestion of model garbage with ammonia stripping. J Biosci Bioeng 111(3):312–319
Yadav M, Paritosh K, Pareek N, Vivekanand V (2019) Coupled treatment of lignocellulosic
agricultural residues for augmented biomethanation. J Clean Prod 213:75–88
184 K. Paritosh et al.
Yan Z, Song Z, Li D, Yuan Y, Liu X, Zheng T (2015) The effects of initial substrate concentration,
C/N ratio, and temperature on solid-state anaerobic digestion from composting rice straw.
Bioresour Technol 177:266–273
Yang L, Li Y (2014) Anaerobic digestion of giant reed for methane production. Bioresour Technol
171:233–239
Yang L, Xu F, Ge X, Li Y (2015) Challenges and strategies for solid-state anaerobic digestion of
lignocellulosic biomass. Renew Sustain Energy Rev 44:824–834
Yao Y, Chen S, Kafle GK (2017) Importance of “weak-base” poplar wastes to process performance
and methane yield in solid-state anaerobic digestion. J Environ Manage 193:423–429
Zhao J, Zheng Y, Li Y (2014) Fungal pretreatment of yard trimmings for enhancement of methane
yield from solid-state anaerobic digestion. Bioresour Technol 156:176–181
Zhu J, Wan C, Li Y (2010) Enhanced solid-state anaerobic digestion of corn stover by alkaline
pretreatment. Bioresour Technol 101(19):7523–7528
Zhu J, Zheng Y, Xu F, Li Y (2014) Solid-state anaerobic co-digestion of hay and soybean
processing waste for biogas production. Bioresour Technol 154:240–247
Zhu Q, Li X, Li G, Li J, Li C, Che L, Zhang L (2020) Enhanced bioenergy production in rural areas:
synthetic urine as a pre-treatment for dry anaerobic fermentation of wheat straw. J Clean Prod
260:121164
Chapter 8
Food Waste Biorefinery for Bioenergy
and Value Added Products
Abstract Food loss and waste (FLW) is becoming a general environmental and
societal problem as well as an opportunity for its valorisation to a plethora of energy
vectors, chemicals and bio-based materials. Food loss is related to the primary and
industrial sectors (i.e., farms and fish farms, factories), while food waste is produced
by retailers and consumers. This leads not only to direct FLW but also indirect loss
of energy and resources devoted to food production. While societal and political
awareness is rising, with the subsequent actions resulting in an efficiency boost
along the food supply chain, unavoidable FLW amounting to more than 1000 Mtons/
year exists due to personal preferences, safety issues and supply inefficiencies.
Likewise, huge amounts of plant biomass by-products (pomace, bagasse, straw,
stover, peels and pulp) over 5000 Mtons/year are generated. First, second and
third-generation biorefineries can be built based on such biomass as well as that
from forest, cattle, fish and algae. Biorefineries are based on thermal, physical,
chemical and biological treatments and can produce a great variety of energy
vectors, namely hydrogen, biogas, bioethanol, biokerosene, biodiesel and biochar;
chemicals (similar to petrochemicals), materials (biomonomers and biopolymers)
and energy (heat). Even feed and food products could be considered as biorefinery
products, ultimately.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 185
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_8
186 M. Ladero et al.
8.1 Introduction
Food Losses and Food Waste (FLW) are considered a major burden to achieve
Sustainable Development on a global basis. Together with other inefficiencies in the
production, distribution and consumption chains of other basic commodities, such as
energy and water, FLW affect the so-called food-energy-water nexus (Slorach et al.
2020). Process efficacy in the production and distribution of food, drinking water
and energy must be promoted to achieve several of the Sustainable Development
Goals (SDGs) included in the United Nations (UN) Agenda 2030. In particular, one
can think of the following SDGs as the most directly affected goals: Zero Hunger (2),
Good Health and Well-Being (3), Clean Water and Sanitization (5), Affordable and
Clean Energy (7), Decent Work and Economic Growth (8), Build resilient infra-
structure, promote sustainable industrialization and foster innovation (9), Ensure
sustainable consumption and production patterns (12) and Conserve and sustainably
use the oceans, seas and marine resources (14). However, all 17 UN SDGs are
affected by the non-optimal extraction, distribution and consumption of these critical
basic commodities (United Nations 2021).
According to the Food and Agriculture Organization (FAO), about 1300 Mtons
of food was wasted or lost in 2013, with an impact on the global economy above US
$750 million (Xue et al. 2017). In fact, when considering the UN SDG 12, by 2030
Responsible Production and Consumption strategies aim to half the current FLW
(Target 12.3) by increasing loss/waste prevention, reduction, recycling and reuse
(Target 12.5). In this sense, there are several indicators proposed and monitored by
FAO. In particular, indicator 12.3.1. is named “Global Food Loss and Waste” and
focuses on FL from production to retail level, but not including the latter
(subindicator 12.3.1.a) and on FW at the retail and consumer levels (subindicator
12.3.1.b). While the former is being applied through 10 main food commodities by
country and annual period, the latter is less amenable to monitoring and only a
proposal for measurement has been created. If we look at the first subindicator,
worldwide FL in 2016 almost reached 14%, with values as low as 5.8% (FAO 2021).
From this perspective, with FL reaching almost 16% of produced food, Europe and
North American countries are far from the less than 6% FL attained in Australia and
New Zealand. FL in the Asia-Pacific region were, in general, low in 2016, less than
8%, which points to a relatively efficient food supply chain to retailer level.
However, these figures could be misleading, most probably due to lack of monitor-
ing or rapid changes in food consumption (India, China). Ample efforts have been
taken in industrialized countries by international agencies (WRAP, FAO, World
Resources Institute and more) to promote adequate monitoring of FLW, defining a
first global standard as recently as June 2016. Nevertheless, several major pitfalls
remain, indicating that global FLW is notably underestimated, even if it is highly
probable than FLW figures for Europe and Northern America are more reliable (Xue
et al. 2017).
The trends of the World population, GDP per inhabitant, total crop production,
and consumption of fossil and biomass-based fuels are compiled in Fig. 8.1a. In
8 Food Waste Biorefinery for Bioenergy and Value Added Products 187
Fig. 8.1 (a) Several key parameters regarding the World population, economic growth and
resource production in the 2000–2019 period (FAOSTAT 2021; World Bank 2021, IEA 2021);
(b) Production of critical crops from 2000 to 2019 (FAOSTAT 2021)
188 M. Ladero et al.
addition, Fig. 8.1b shows the worldwide trends in the main crops during four 5-year
periods starting in 2000. Although the steady linear growth in the World population
results in an increase of all other variables, it is evident that trends are relatively
different. Some are linear, such as population and crops, whilst other are asymptotic
(GDP/cap; fossil fuels). A closer look at the evolution of different crops (Fig. 8.1b)
indicates that cereal crops, the most abundant, strongly increase (38% growth in the
2000–2019 period), while roots and tubers only experience a growth slightly higher
than 20%. On the contrary, oil crops dramatically increase by 87% in the mentioned
20-year period. Thus, crops maintain the same linear trend as the World population,
but with an increase of 43% in comparison to 19.78%, respectively. This leads to an
economic growth and an increase in the general well-being, as shown by GDP per
inhabitant increments, though the latter seems to be reaching a stable value around
US$12,000 (2010 value) per inhabitant and year. The last 20 years have undergone a
booming hyperbolic economic growth, much in line with the progressive reduction
in consumption of fossil resources. By 2000, the World seemed to reach another
stable value of about US$5400 (2010 value)/cap/year. In this period, there were
important political and economic paradigm shifts (Euro introduction, Gulf crisis,
globalization), which led to a dramatic increment in global wealth growth. As we
reach another plateau, it seems evident that new paradigms are to arrive, possibly
including concepts such as sustainable growth, circular economy, bioeconomy, big
data analysis and industry 4.0 (Dantas et al. 2021). If we considered the exponential
growth in renewable electricity sources (wind and solar) [378% from 2000 to 2018,
reaching a 2% share of the global energy consumption] and the steady increment in
biomass resources for energy, these are called to complement and, ultimately,
substitute fossil energy resources. Notwithstanding, considering an increase of
44% and 52% in gas and coal production in the last 20 years, substitution will
need political action, social perception changes and dramatic technological advance-
ment on energy production, distribution and consumption (LLNL 2021).
European consciousness on the problem posed by food waste (FW) is reflected in
the shift of municipal solid waste (MSW) fate in the last decade. MSW contains up to
50% FW and, by 2005, was mainly disposed of in landfills, while, by 2015, >25% of
it is recycled, with the EU objective to get to 60% in 2025 (Carmona-Cabello et al.
2018). In this valorization strategy, the development of second generation
biorefineries is a must, not only to cope with an important, though, hopefully,
decreasing amount of FW in the future, but also to profit from biomass created
during the food production activities in the primary sector.
Thus, biorefineries are intended to be relevant in an industrial ecology strategy
focused on the use of renewable resources (biomass) through closely integrated
processes in the fashion of existing refinery processes to the plethora of chemicals,
materials, fuels, food and feed needed by mankind. In this sense, both bioeconomy
and circular economy can be established with the aid, among others, of tools such as
life cycle assessment (LCA) and life cycle cost (LCC) in the design of processes and
products (Lam et al. 2018). An integration of human activities in nature, if adequate,
8 Food Waste Biorefinery for Bioenergy and Value Added Products 189
Fig. 8.2 Biorefinery approaches to FLW and food-related biomass valorisation: The thermochem-
ical and the fractionation approximations
can tackle complex problems such as global climate change, resource scarcity and
increasing demands due to growing population and global living standards (Mohan
et al. 2019).
In biorefinery schemes, the chemical composition and physical structure of any
biomass feedstock should be considered to establish its potential for the generation
of products and to obtain positive synergism with other resources of similar or
complementary nature (Carmona-Cabello et al. 2018), including fossil resources
(for example, in the production of hydrotreated vegetable oil (HVO)). Likewise, the
water content can drive towards the need of drying or the selection of fermentation or
hydrothermal processes instead of pure thermal or catalytic ones. In such cases
pyrolysis or other operations common to the thermochemical biorefinery (Kim
et al. 2020) could be undertaken. Additionally, the pretreatment-saccharification-
transformation scheme typical of the fractionation bio-chemical biorefinery (Rastogi
and Shrivastava 2017) could be approached (Fig. 8.2).
In the fractionation biorefinery, carbohydrate-rich biomass has been typically
addressed by a series of pretreatments, saccharification and/or transformation by
chemical and biological means to bioethanol, biokerosene and biodiesel (as a mix-
ture of fatty acid methyl ester (FAME)). Since food-designed resources have been
firstly approached due to their reactivity/accessibility, first-generation biorefineries
have encountered some problems due to the food/fuel controversy. Therefore, due to
its greater availability and non-edible nature, lignocellulosic and algal biomass is
being considered of high interest as feedstock. Lignocellulosic biomass is well-
known for its recalcitrant nature, mainly related to its lignin content. Hence,
190 M. Ladero et al.
The largest bioethanol producing country is the United States, with around 55%
of world production, followed by Brazil (30%), the European Union (5%), China,
Canada, India, Thailand and Argentina. The main expansion of this biofuel is
expected to take place in developing countries, with Brazil, India and Thailand,
together with China, being the most important (OECD/FAO 2018).
However, the feedstock currently used for its production is corn, sugarcane,
barley, wheat, sugar beet and sweet sorghum; agricultural products used for food,
whose use in obtaining bioethanol is framed within first generation biofuels (Cooper
et al. 2020). First generation biorefineries are based on raw materials usually
employed for food and feed, very rich in simple sugars (monosaccharides, disac-
charides) or edible polysaccharides (as starch). Therefore, as food production, in
term of mass, is far from the production of fuels need by Humanity (several billions
of tons annually), a high stress is created on food market, this being the main reason
for the search for other raw materials. A main advantage of first-generation ethanol
production is the process simplicity—very well-known due to its long-time use in
the food industry. Although the cost of biomass purchasing is high, up to 65% total
costs, lower technology and process-related costs in comparison to other raw
materials makes first-generation bioethanol production the first and most developed
option nowadays. As for the use of non-edible lignocellulose-based feedstock
(second-generation biorefineries), it involves 20–25% higher costs due to the tech-
nical complexity of required pretreatment and saccharification steps and the higher
complexity of the fermentation step, related to the antimicrobial activity present in
hydrolysates (Esmaeili et al., 2020). Thus, tax policies and monetary incentives are
needed for the transition from first to second generation biorefineries.
Fermentation varies depending on the raw material used, but most of its stages are
maintained with certain changes in process variables, such as pH (between 3.7 and
5.5) or temperature (between 30 and 40 C). Milling, fermentation and distillation
are performed in-series when using sugarcane and other sources of soluble sugars. It
is necessary to add gelatinization and hydrolysis stages (enzyme-based mashing-
saccharification) prior to fermentation when polysaccharide-rich feedstock is
employed (Ayodele et al. 2020). In both cases, up to 95% glucose or sucrose present
in the feedstock can be obtained. Fermentation is usually carried out with Saccha-
romyces cerevisiae, having genetically modified yeasts proven to enhance its pro-
ductivity. Finally, in a multistage distillation process, 95% ethanol is obtained,
which can be subsequently dehydrated to 99.99% using azeotropic distillation
(Ayodele et al. 2020) or membrane technology (Khalid et al. 2019). This technology
has been used in the United States, Brazil and the EU for more than two decades.
First-generation bioethanol production competes for farmland with food produc-
tion. This issue is at the centre of sustainability debates as large areas of arable land
with a high need for rainfall or irrigation are required (Ayodele et al. 2020). Thus,
most recent research focuses on obtaining the so-called second-generation
(2G) bioethanol. The raw materials used to produce 2G ethanol are classified into
three main groups: agricultural residues, forestry residues and energy crops. This
chapter puts the spotlight on agricultural and agrofood residues due to their abun-
dance, particularly sweet sorghum and sugar beet bagasse, and straw and corn
192 M. Ladero et al.
stubble. The chemical composition of these residues presents some variation in their
components: cellulose (between 33 and 48%), hemicellulose (between 18 and 27%)
and lignin (between 7 and 21%) (Ayodele et al. 2020; Jeevan Kumar et al. 2020).
Lignocellulose pretreatments are more complex than those applied to starch or
feedstock rich in free simple sugars (first-generation biorefineries). These pretreat-
ments aim at obtaining cellulose, hemicellulose and lignin fractions with adequate
purity to boost the development of 2G ethanol (Romaní et al. 2018; Cheah et al.
2020). They are critical to the economics of 2G ethanol.
The pretreatments studied are very varied and can be classified into different
categories: physical processes (milling, extrusion, freezing, microwave and ultra-
sound processes), chemical processes (treatment with acids, bases or ionic liquids,
organosolv, oxidation with sulfite, alkaline wet oxidation and ozone), physicochem-
ical processes (steam explosion, treatment with liquid hot water and ammonia fibre
explosion) and, finally, biological processes (digestion by fungi and bacteria) (Cheah
et al. 2020; Dhiman and Mukherjee 2020; Jeevan Kumar et al. 2020).
Hydrothermal treatments enclose the use of supercritical water as a quite
established operation, which requires temperatures higher than 374 C. This condi-
tion provokes both the denaturalization of the walls of plant cells, the degradation of
hemicellulose and the transformation of lignin into its monomers. The advantages of
this pretreatment are high energy conversion, low corrosion and no need for catalysts
(Ayodele et al. 2020). There are some problems related to the high temperature and
pressure values required, entailing the use of sophisticated equipment, special
materials and reducing gases. Hydrothermal treatment includes emerging technolo-
gies as the use of ionic liquids. Ionic liquids modify the lignocellulose biomass
structure in a way that requires subsequent stages of regeneration, filtration, washing
and distillation prior to hydrolysis and fermentation (Ayodele et al. 2020).
An effective pretreatment must provide large amounts of cellulose and hemicel-
lulose prior to hydrolysis to monomers (mainly glucose and xylose). In addition,
pretreatment causes the generation of inhibitory compounds for subsequent stages
(enzymatic hydrolysis, if applicable, and fermentation). The pretreatments that
generate highly inhibiting compounds are acid pretreatment, organosolv and steam
explosion (Cheah et al. 2020).
The following hydrolytic stage is an enzymatic process that can be combined with
the fermentation stage (Jeevan Kumar et al. 2020). The basic operation uses two
separate stages and is known as ‘Separate Hydrolysis and Fermentation’ (SHF). Its
main disadvantage is the inhibition of cellulolytic enzymes by the accumulation of
reducing sugars in the reaction medium during the enzymatic stage. Also, since two
reactors are needed, the investment in process equipment becomes more expensive
(Jeevan Kumar et al. 2020). To solve the problems posed in the SHF approach, the
‘Simultaneous Saccharification and Fermentation’ (SSF) process was developed, in
which the hydrolysis and fermentation stages are carried out jointly in a single
reactor. Obviously, this procedure is possible only with similar optima conditions
for both transformations (Jeevan Kumar et al. 2020). Another advancement is
‘Consolidated Bioprocessing’ (CBP), which consists of simultaneous enzyme pro-
duction, saccharification and fermentation by the same microorganism, decreasing
8 Food Waste Biorefinery for Bioenergy and Value Added Products 193
the need for exogenous enzymes and reducing operation costs (Jeevan Kumar et al.
2020). A last alternative is the partially consolidated bioprocess (PCBP), a combi-
nation of simultaneous pretreatment and saccharification and subsequent fermenta-
tion (Jeevan Kumar et al. 2020).
Pretreatment and saccharification stages can jointly result in a complex mixture of
hexoses (glucose, mannose and galactose) and pentoses (xylose and arabinose).
Therefore, microorganisms (as pure or mixed cultures) capable of transforming all
of them to ethanol whilst being highly ethanol tolerant are needed (Ayodele et al.
2020; Jeevan Kumar et al. 2020). S. cerevisiae is the most widely used microorgan-
ism in the production of 1G bioethanol from glucose. However, 2G bioethanol
production faces the described challenges, requiring other microorganisms in addi-
tion (Ayodele et al. 2020; Jeevan Kumar et al. 2020) or to obtain ad hoc cell factories
to produce bioethanol using different approaches such as metabolic engineering,
genetic engineering or adaptive evolution (Sakar et al. 2020; Sharma and Arora
2020).
One strategy is the use of ethanol-producing bacteria (EPB), such as Zymomonas
mobilis, capable of fermenting both types of sugars. Another strategy is based on the
use of mixed cultures, called co-fermentation. In this case, microorganisms capable
of fermenting pentoses, such as Candida shehatae, Kluyveromyces marxianus,
Pichia stipis and Pachysolen tannophilus, are used together with S. cerevisiae. In
the case of using co-fermentation, the SHF and SSF processes (SHF, SSF) are
referred to as ‘separate hydrolysis and co-fermentation’ (SHCF) and ‘simultaneous
hydrolysis and co-fermentation’ (SSCF) (Ayodele et al. 2020; Jeevan Kumar
et al. 2020). Finally, the combined production of 1G and 2G bioethanol can lead
to profitable processes (Ayodele et al. 2020).
As for starch-rich FW, although fermentation approaches are similar to those
addressed above, the enzymatic saccharification is achieved by amyloglucosidases,
α-amylases, fungal mixtures, including proteases for more complex food matrixes
still rich in carbohydrates (Saeed et al. 2018). In fact, the treatment of household FW
benefits from the use of complex mixing systems within the reactors and the use of
several enzymatic cocktails (Loizidou et al. 2017). Finally, ethanol fermentation is a
valuable pretreatment in the process of biogas (or methane) production by anaerobic
fermentation (Sun et al. 2020; Karimi and Karimi 2018).
8.2.2 Biobutanol
It also shows easy transportation and less corrosion because of its low vapour
pressure, has a vaporization heat slightly higher than gasoline and its low solubility
in water reduces the potential of groundwater contamination (Pugazhendhi et al.
2019; Bankar et al. 2013). A summary of the different processes to produce
n-butanol and isobutanol is shown in Table 8.1.
Although bio-n-butanol had been used on a large scale and its biologically
fermentative production has a long history, the higher octane number of isobutanol
8 Food Waste Biorefinery for Bioenergy and Value Added Products 195
has attracted increasing attention in the fuel industry. However, biological isobutanol
is harder to separate and purify due to its low concentration in the fermentation broth
(Fu et al. 2020; Bankar et al. 2013).
In the petrochemical route, n-butanol can be produced using three industrial
processes: propylene oxo synthesis, conversion of n-butanol from ethanol and the
Reppe synthesis, oxo synthesis being the most common in industry. In this process,
propylene and syngas feed streams are converted with cobalt or rhodium catalysts
(Huzir et al. 2018; Jin et al. 2011).
n-Butanol serves as a precursor for producing paints and plastics and is also an
intermediate in the production of butyl acrylate, methacrylate and as a solvent in the
production of dyes, oils and waxes. The n-butanol market size is forecasted to reach
US$6.74 billion by 2025 with an estimated annual growth rate of 6.5% during
2020–2025. Major players in the n-Butanol market are Arkema (France), BASF
(Germany), Dow Chemical (US), Eastman Chemical (US) and Mitsubishi Chemical
(Japan) (The Express Wire 2020). However, butanol and its isomers present the
disadvantage of a quite low production compared with ethanol fermentation, which
is 10–30 times higher (Bankar et al. 2013; Chen et al. 2014).
On the other hand, the global isobutanol market is expected to reach US$1.18
billion by 2022. In 2014, synthetic isobutanol accounted for 58% of the total market,
while bio-based isobutanol is expected to grow 7.0% from 2015 to 2022. Some of
the main companies involved with biobutanol production are Butamax (US), Gevo
(US), BUTALCO (Switzerland), Green Biologics (UK) and Cathay (China) (Grand
View Research 2016).
Agricultural and food residues with a high content of cellulose and hemicellulose
and low lignin content are an ideal substrate for biobutanol production. This
production occurs in two different steps, mainly extracellular breakdown of poly-
meric carbohydrates of biomass and intracellular two-stage ABE fermentation using
C. acetobutylicum. In the first step, hydrolysis of the polymers is made by the
numerous enzymes secreted, while in the ABE fermentation acetone, butanol and
ethanol are produced in 3:6:1 ratio (Pugazhendhi et al. 2019; Bankar et al. 2013;
Chen et al. 2014).
Butanol toxicity due to its accumulation is the major problem during fermenta-
tion. Producing >7.4 g/L of butanol suppresses Clostridium growth and limits
butanol yield leading to an increase of the production and recovery costs. To solve
this issue, there is an ample toolbox in genetic engineering of the biosynthesis
pathways. Thus, researchers have developed solventogenic Clostridia tolerant
strains towards butanol production and have restricted their spore formation.
Besides, interest has also been shown in E. coli and S. cerevisiae, which are well
tractable and rapidly growing (Pugazhendhi et al. 2019; Fu et al. 2020; Bankar et al.
2013)
Batch bioprocessing is the most used method to conduct the fermentation,
although, in an economical comparative investigation, continuous fermentation
was found to be preferred for large-scale production. Among the techniques to
recover butanol from the fermentation broth, distillation is the most traditional and
widely applied method. However, due to water having a lower boiling point than
196 M. Ladero et al.
butanol, researchers have developed new recovery techniques to decrease the cost of
the separation, like adsorption, gas stripping and pervaporation (Pugazhendhi et al.
2019; Bankar et al. 2013).
Due to the problems associated with the ABE fermentation and the increasing
urgency to substitute petrochemical routes, ethanol upgrading to butanol has
received much attention in recent decades. There are two main pathways to achieve
this transformation, namely: the direct dimerization of two molecules of ethanol and
the multistep tandem synthetic route known as the Guerbet reaction. The later
includes the dehydrogenation of ethanol, followed by aldol condensation (under a
strong inorganic base like NaOEt), dehydration and final hydrogenation to produce
n-butanol or isobutanol (Wu et al. 2018; Kulkarni et al. 2018; Wingad et al. 2016).
To carry out this upgrading, recent progress in catalyst development has been made,
which involves homogeneous catalysts, such as iridium and ruthenium complexes;
and heterogeneous catalysts, including metal oxides and zeolites like MgO or Al2O3,
hydroxyapatite (HAP) and supported metal catalysts like Ni/Al2O3 have been
explored. These reactions are performed under relatively high temperatures and
pressures (Wu et al. 2018; Kulkarni et al. 2018; Wingad et al. 2016).
8.2.3 Biodiesel
Biodiesel is defined as the monoalkylesters of long chain fatty acids derived from
renewable liquid feedstocks for use in compression ignition engines. Biodiesel is a
worldwide recognized potential alternative for substituting petroleum-based diesel.
Biodiesel has a series of advantages compared with conventional diesel, including its
renewable origin, biodegradability, non-toxicity and the virtual nonexistence of
sulfur and aromatics compared to diesel. In addition, biodiesel is of geostrategic
value for countries as it provides support to local economies and reduces energy
dependency (Hajjari et al. 2017; Foroutan et al. 2020; Hamza et al. 2020).
The key aspects of biodiesel application are the choice of feedstock and the
design of a suitable production process. The main drawbacks for the implementation
of biodiesel are the production costs, associated to raw material costs, and their
effective supply as well as the development of efficient production processes to
satisfy the demands as fuel. The properties of biodiesel as fuel and the required
process production (including previous treatment and purification) development are
strongly correlated with the fatty acid content or degree of unsaturation of the
feedstock. Conventionally, biofuels including biodiesel are classified based on
their feedstock and production technologies into different generations. These gen-
erations include biofuels produced from edible oil seeds, non-food oil crops and
wastes, algae, and the genetically engineered oil crops. The debate between using
edible and non-edible oil feedstock for biodiesel is driven by feedstock availability
and cost, greenhouse gas emission, economic efficiency of using fertile lands as well
as fuel vs. food/feed competition. A suitable alternative for biodiesel relies on
non-edible vegetable oil; however, controversy arises regarding the use of soil
8 Food Waste Biorefinery for Bioenergy and Value Added Products 197
resources for the massively needed supply. In this context, waste-oriented oils/fats
have been proposed as the excellent options to produce biodiesel (Ramos et al. 2019)
with a major dual advantage: cost and environmental impact. Utilizing it as feed-
stock for biodiesel means a double contribution in making the environment safe and
cleaner. Hence, waste based-oil is considered an economically and socially viable
feedstock for low cost biodiesel production on a commercial scale (Hamza et al.
2020; Goh et al. 2020).
Biodiesel is synthetized by the process of esterification or transesterification
reactions of fatty acid or vegetable oil in the presence of monohydric alcohol,
catalyst, and defined conditions of temperature over a period of time.
Transesterification is the catalysed process of trading the alkoxy group of an ester
by an alcohol such as methanol and ethanol (acyl acceptor) to convert the triglyc-
erides of the oil to fatty acid alkyl esters and glycerol. In biodiesel production,
different alcohols can be used. Methanol and ethanol, and the reaction product
produced when methanol is used is called a fatty acid methyl ester mixture
(FAME), whereas if the alcohol is ethanol, the product obtained is a fatty acid
ethyl ester mixture (FAEE). The mild reaction conditions needed, the fast reaction
time and the easy phase separation combined with its low-cost and industrial
availability make methanol the most used alcohol in biodiesel production, although
it is not a renewable bioresource as ethanol (Ramos et al. 2019).
Besides the properties of the original feedstock, the type of catalysts used in the
transesterification reactions is another critical element affecting the biodiesel pro-
duction. They could be chemical compounds, such as acids and/or bases, and
enzymes, working in homogenous or heterogeneous phase. The overall process
involves three consecutive reversible reactions which produce intermediate mole-
cules of di- and monoglycerides. Reactions involved and catalysts have been
reviewed in detailed in recent literature (Ramos et al. 2019; Changmai et al. 2020;
Hamza et al. 2020). The following paragraphs cover the basics of reaction catalysis
in the interplay with feedstock properties.
Homogeneous base catalysis was vastly implemented due to its simplicity, mild
reaction conditions and high productivity and conversion. However, it is very much
limited by the free fatty acid content. The homogeneous acid-catalysed reaction is an
alternative since it can simultaneously catalyse esterification and transesterification.
However, it deals with a slower reaction rate and the energy and methanol demand is
higher. As a consequence, general drawbacks of the homogeneous catalysis are the
result of side reactions of saponification and hydrolysis; or the high capital cost and
energy required for the process, the high cost of separation and purification of
catalysts and glycerol as well as the need for neutralization and wastewater treatment
(Ramos et al. 2019).
To overcome the previous disadvantages, heterogeneous catalysed reactions have
been developed involving different solid materials. Heterogeneous catalysts facili-
tates separation and are less sensitive to high FFA content. As drawbacks, they are
costly and energy intensive as high temperatures and high alcohol:substrate ratios
are needed (Ramos et al. 2019).
198 M. Ladero et al.
reaction mechanisms and conditions, the emphasis is to improve the blending of the
solvents to enhance the rate of reaction and produce higher biodiesel yields
(Kirubakaran and Arul Mozhi Selvan 2018; Hamza et al. 2020; Hamza et al. 2020).
Before commercialization, some purification procedures may be performed to
maintain quality standards for the final product to be considered biodiesel. Impurities
come from solids, unsaponifiable materials present in the feedstock itself, including
catalyst residues, water, glycerol and excess alcohol from the reaction, such as polar
compounds, dimers, mono and diacylglycerols and free fatty acids. There are several
methods for removing such impurities, such as membrane filtration, centrifugation,
and distillation as well as wet and dry washing (Fonseca et al. 2019; Ramos
et al. 2019).
Last trends of manufacturing biodiesel from waste oil focuses on developing new
heterogeneous catalysts, enabling technologies, reactor concepts (Gupta et al. 2020)
and reaction media to intensify the reactions and increase productivity (Bankovi-
ć-Ilić et al. 2017; Degfie et al. 2019; Ching-Velasquez et al. 2020). New heteroge-
neous chemical catalysts include waste egg shell and heterogeneous base catalysts or
carbon based nanocatalysts (Banković-Ilić et al. 2017). Immobilized lipases as
heterogeneous biocatalysts are under constant development to increase productivity
and decrease production costs (Cavalcante et al. 2021). Recent examples are the
application of lipases from Rhizomucor miehei (RML), lipase B from Candida
antarctica (CALB) and lipase from Thermomyces lanuginosus to overcome current
limits of enzyme catalysis to increase the biodiesel production yield and rate (Babaki
et al. 2017; Badoei-dalfard et al. 2019; Ching-Velasquez et al. 2020). Intensification
of the reaction is being assisted by enabling technologies as microwave assisted
intensification of biodiesel production from waste cooking oil using heterogeneous
base catalysts (Gupta and Rathod 2018) or ultrasound system for the enzymatic
transesterification of oils using combi-lipases as biocatalyst (Poppe et al. 2018).
Among new reaction media, the use of low-cost deep eutectic solvents (DESs) has
been investigated as a new reaction medium for enzymatic biodiesel production from
waste oils with reuse of the catalyst (Merza et al. 2018).
The fermentation of organic substrates, including FW, can lead to gaseous products
of interest mainly in the energy sector, but also as building blocks for many other
industries. Biogas is one of such products, which is defined as a mixture of methane,
CO2 and small amounts of other gases produced by anaerobic digestion (AD) of
organic matter. Commonly, the methane content is in the range from 45 to 75% v/v,
with most of the remainder being CO2. The composition varies with the feedstock
and the fermentation process used, which will inevitably affect the energy content of
biogas, whose lower heating value (LHV) varies from 16 to 28 MJ/m3. Globally, the
200 M. Ladero et al.
production of biogas has steadily increased over the last two decades from 0.28 EJ in
2000 to 1.33 EJ in 2017, with the EU contributing as much as half of the supply,
followed by Asia with about a third. However, as of 2017, this is only about 2% of
the overall bioenergy supply and has potential to represent a higher share in
estimated growing capacities that only in the EU could increase up to 20 109 m3
(World Bioenergy Association 2019).
Biomethane, also referred to as renewable natural gas, consists mainly of pure
methane, which can be obtained by purification and upgrading of biogas by remov-
ing CO2 and other gases or by gasification of solid biomass followed by methana-
tion. It has a LHV of 36 MJ/m3 and can use the distribution channels and be applied
in technologies that currently use natural gas. The biomethane market prospects also
show an outstanding potential increase in demand. The International Energy Asso-
ciation estimates growths up to about 45 Mtoe by 2030 and 75 Mtoe under a Stated
Policies Scenario, which could reach as much as 110 Mtoe by 2030 and above
200 Mtoe by 2040 under a Sustainable Development Scenario with developing
countries in Asia leading the way. This product sees a broad-based growth across
all sectors where natural gas is present (buildings, power and heat, industry), whilst
also tapping into markets like transport (International Energy Association 2021).
Hydrogen shows great advantages as fuel candidate, as it has a high energy yield
(142.4 kJ/g) and heating efficiency (about 2.75-fold with respect to hydrocarbons). It
is a clean fuel as only water is produced from combustion and can made use the
infrastructure associated with the rapid development of fuel cell technologies (Wang
and Yin 2018). Current worldwide hydrogen value chains currently indicate a
production of hydrogen from fossil-based natural gas and coal amounting to
196 Mtoe and 75 Mtoe, respectively. The production of hydrogen as by-product of
other processes is about 48 Mt H2, only 0.3 of which comes from renewables, which
points towards an opportunity of increasing the share. By 2050 transport, industrial
energy and particularly the use of hydrogen as industrial feedstock will have an
increased market share, such as in methanol and ammonia production (International
Energy Association 2021). As for biohydrogen in particular, it was calculated that
about 15.5 Mt could be produced using waste from crops like potatoes, wheat, corn,
sugarcane or barley by 2030 (Alavijeh et al. 2020).
In connection with methane and hydrogen, the term hythane was coined and
patented by the company Hydrogen Component Inc., which tested mixtures of these
gases in internal combustion engines. Mixtures containing 20% v/v H2 showed a
reduction of NOx emissions whilst maintaining energy efficiencies. Significantly,
these mixtures did not require changes in engines or infrastructures.
Finally, the term volatile fatty acids (VFA) typically refers to organic acids of
chain length from two to up to six carbons. VFAs have been used in many
applications, including energy related technologies, such as the generation of elec-
tricity in microbial fuel cells, in the synthesis of complex polymers or in textiles,
food and pharma products. In general, the market value of VFAs is related to the
length of the chain, which drives the production towards heavier products rather than
lighter, there is even a trend towards the generation of higher fatty acids (Esteban and
Ladero 2018).
8 Food Waste Biorefinery for Bioenergy and Value Added Products 201
Policies to promote the generation of all of these gaseous products from renew-
able substrates, particularly FW and other agricultural wastes show a particular
alignment with the UN SDGs, particularly SDG 7 (Affordable and clean energy)
and SDG 12 (Responsible consumption and production). The drive towards the
production of each of the mentioned products from organic substrates will greatly
depend on the fermentation technology used. The following epigraphs will cover the
particularities of their production with a focus on FW. Table 8.2 summarizes the
main results of selected references of processes to obtain the mentioned products.
Biogas is produced by single-stage AD of organic matter such as wastewater
sludge, manure, crop residues or, naturally, FW. AD is a process that consists of four
steps, namely: (1) hydrolysis, where lipids, proteins, carbohydrates and other large
polymers are broken down into smaller molecules by anaerobic bacterial consortia
including Cellulomonas clostridium, Bacillus thermomonospora, Ruminococcus
baceriodes among others; (2) acidogenesis, during which bacteria like Streptococ-
cus, Lactobacillus or Bacillus transform the monomers into short chain organic acids
and alcohols; (3) acetogenesis, where the former products are converted into H2,
CO2 and acetic acid by bacteria like Clostridium, Acetobacterium or
Syntrophomonas; and (4) methanogenesis, in which Methanolobus, Methanosarcina
or Methanococcus reduce the previous products to CH4.
The first AD plant treating FW dates from 1939 in the USA, although this type of
processes has attracted interest in Europe in the last decades, reaching a capacity of
150 Mton/year (Bolzonella et al. 2018). The quantity and quality of biogas, i.e., its
biomethane content, depend on the type of inoculum and operational factors like pH,
temperature, hydraulic retention time (HRT), solid retention time (SRT), stirring and
loading rate (Tabatabaei et al. 2020). Largely, the biogas produced is also influenced
by the composition of the feedstock. In the case of FW, this can be disadvantageous
as they could have low C/N ratios, high lipid concentrations, insufficient macro- and
micronutrients or there could be toxic compounds, all of which may trigger different
inhibition mechanisms that may unfavour the steps of AD to different degrees
(Tabatabaei et al. 2020).
Biogas production is a slow process owing to the long doubling time of the
microorganisms involved in these stages, being methanogenesis the rate-limiting
stage in the AD process leading to higher HRTs. In other cases, the hydrolysis stage
can be cumbersome if complex organic substrates are used, hence leading to toxic
byproducts and VFAs. To decrease HRTs and enhance biogas production, one
possible solution involves the upstream pretreatment of the feedstock to increase
its digestibility. In this sense, biological pretreatment using fungi, microbial consor-
tia or enzymes is beneficial as they are relatively simple, do not require additional
chemicals and the investment is typically low. On the other hand, such pretreatments
require long reaction times and may be hard to apply in some substrates. Further
lines of work on pretreatment feature techniques like microaeration and metabolic
engineering of microorganisms (Tabatabaei et al. 2020). Additionally, components
like fats or terpenes need to be removed to enhance the productivity as they show
inhibition effects and methane formation can be reduced. In the case of terpenes,
these compounds are known to have antimicrobial and antioxidant properties and are
202
Table 8.2 Examples of biological processes to obtain biogas, biohydrogen, biohythane and VFA from FW sources
Food waste Type of AD Inoculum Pretreatment reactor Operating conditions Main results Reference
Fruit, vegeta- Two-stage Activated sludge Mechanical homogeniza- T ¼ 35 C YCH4 ¼ 460 mL/g Wang et al.
ble and (acidogenic and from wastewater tion HRT ¼ 10 days + 20 VS (2014)
kitchen waste methanogenic) treatment plant Pilot plant CSTR (2 and days CVFA ¼ 2582 mg/
4 m3) OLR ¼ 4.5 g VS/ (L day) L
Tomato Mainly Methanogenic sludge Lycopene extraction and T ¼ 55 C YCH4 ¼ 105 mL/g Allison and
pomace methanogenic from UC Davis enzymatic saccharifica- HRT ¼ 90 days VS Simmons
tion CSTR OLR ¼ 1 g VS/ (L day) (2017)
Kitchen waste Full four-stage oper- Activated sludge Lipid extraction T ¼ 55 C YCH4 ¼ 531 mL/g Algapani
ation codigested with from wastewater Shaken bottles (120 mL) HRT ¼ 30 days VS et al. (2017)
sludge treatment plant OLR ¼ 4.24 g VS/ CVFA ¼ 510 mg/g
(L day) VS
Household Two-stage Brown water from None for FW; preheating For H2: YCH4 ¼ 728 mL/g Paudel et al.
food waste (acidogenic and University toilets for sludge CSTR (10 and T ¼ 37 C; HRT ¼ 0.3 VS (2017)
methanogenic) 40 L) days; OLR ¼ 47.8 g VS/ YH2 ¼ 99.8 mL/g
(L day) VS
For CH4: CVFAb¼ 2200 mg/
T ¼ 37 C; HRT ¼ 20 L
days; OLR ¼ 37.2 g VS/
(L day)
M. Ladero et al.
8 Food Waste Biorefinery for Bioenergy and Value Added Products 203
also used as food colorant, so isolation of this type of compounds could be interest-
ing not only from the perspective of increasing yields of biogas, but also towards an
integral valorisation of FW (Allison and Simmons 2017).
Hydrogen from different biomass sources can be obtained through different
(thermo)chemical methods that include the gasification of hydrocarbons, steam or
aqueous phase reforming and the corresponding water-gas shift reactions. In addi-
tion, a number of biological processes using lignocellulosic and carbohydrate-rich
substrates like FW have attracted interest in the last few years, including direct and
indirect biophotolysis (Sampath et al. 2020). Both require phototrophic microorgan-
isms like cyanobacteria or microalgae like Chlorella or Chlamydomonas to capture
solar radiation and perform oxygenic photosynthesis, by means of which H2 can be
generated in a single or two-step process.
In dark fermentation, bacteria like Clostridium, Enterobacter, Cellulomonas or
Thermotoga can be used, in which hydrogenases are the key catalyst for the
generation of H2, which is produced without any oxygen or light inputs. Feedstock
like FW or other type of waste can be used, although it appears that lipid and protein-
rich substrates are detrimental to H2 production compared to others rich in carbo-
hydrates (Ntaikou et al. 2010).
In photofermentation, light-dependent purple sulfur and, especially, non-sulfur
bacteria like Rhodobacter sphaeroides or Rhodopseudomonas palustris are suitable
to decompose organic acids to H2 with nitrogenases and hydrogenases catalyzing the
process (Hitam and Jalil 2020).
Hybrid systems can increase biohydrogen productivity combining
non-photosynthetic and photosynthetic bacteria. The organic acid byproducts gen-
erated during dark fermentation by bacteria like Clostridium butyricum serve as
substrate for photoautotrophic microorganisms like Rhodopseudomonas faecalis in
the presence of light in the second stage (Azwar et al. 2014).
Apart from physical pretreatments that are also applicable to the production of
biogas to make the complex polymeric structure more accessible in FW or other
biomass, certain chemical pretreatments are meant to increase the yield to hydrogen.
For example, chloroform, BESA and O2 inhibit methanogenic species (Wang and
Yin 2018).
Owing to the importance of hydrogen production processes, a techno-economic
analysis of a dark fermentative hydrogen production process from FW in China
recently showed that it could be profitable operating with a supply of above 0.3 ton/
day of FW. The unit production cost of the product would be US$1.02 per cubic
meter with labour having the highest impact on the costs (Han et al. 2016).
In the past few years, the production of biohythane has attracted a good deal of
attention, for which the use of FW and sewage sludge have had prominent use.
Contrary to the sole production of biogas, the substrates are subjected to a two-stage
AD consisting first of a dark fermentative phase followed by a methanogenic step.
This operation allows the reduction of the overall required fermentation time and
reactor volumes, which makes it attractive from the operation point of view
(Bolzonella et al. 2018). The two main types of feedstock typically used are FW
and sewage sludge or codigestion of both. LCA research on biohythane shows that a
204 M. Ladero et al.
Fig. 8.3 Concept for the production of furans from carbohydrate-rich FW using a biphasic liquid
system
based (ChCl) deep eutectic solvents can help hydrolyse carbohydrate polymers. A
biphasic system with ChCl-H2O as polar phase and GVL as extracting phase was
reported for the production of HMF from carbohydrate-rich FW like bread or rice
waste, for which AlCl36H2O was used as catalyst. Yields up to 18.60% were
obtained with GVL greatly helping to increase the extraction efficiency, thereby
avoiding the degradation of HMF. Very importantly, the ChCl-H2O solvent showed
high recyclability (Ji et al. 2020).
Citric acid (CA) is one of the most widely used organic acids: its production was
2 Mtons in 2020, with China being the world’s largest producer. The global market
could reach US$3.6 billion in 2025, growing more than 5.24% by then. Currently, it
is mainly consumed in the food, pharmaceutical and personal care industries as
acidulant, antioxidant, preservative, flavouring, buffer and astringent. Other appli-
cations include leather tanning, metal finishing and phosphate replacement in deter-
gents (Amato et al. 2020, Mores et al. 2021). Great research efforts have focused on
the production of citric acid through fermentation processes. Aspergillus niger is the
main producer due to its easy manipulation and its high selectivity/yield towards the
acid of interest in comparison to other microorganisms such as Yarrowia lypolytica
yeast, which also generates isocitric acid (Mores et al. 2021). A. niger can produce
CA from glucose, sucrose, maltose, galactose, mannose and fructose. This is a good
reason to employ impure raw materials with high carbohydrate content such as cane
molasses or corn starch.
In recent years, several studies have focused on FW valorisation to reduce costs
(Amato et al. 2020). Ozdal and Kurbanoglu (2019) produced 68.6 g/L citric acid
through the most widely used method in the industry, submerged fermentation. The
best results were observed with 0.15 g/L of KH2PO4 and 4 g/L of chicken feather
peptone (CFP). This nitrogen source was obtained from a complex pretreatment with
two consecutive hydrolysis with 6 N H2SO4 at 70 C, 24 h and 130 C, 4 h, followed
by the addition of different hydroxides, filtration and evaporation. The carbon source
was obtained by pretreatment of sugar beet molasses with H2SO4 1 N at 90 C,
15 min and its subsequent addition of 0.1 g/L of K4Fe(CN)6. Employing the same
type of operation, Liu et al. (2018) used waste cooking oil as a carbon source,
co-producing 12.6 g/L of citric acid and 4 g/L of erythritol at pH 6 and an osmotic
pressure of 0.75 osmol/L. Hou and Bao (2018) produced 136.3 g/L of citric acid by
simultaneous saccharification and fermentation (SSF) of corn stover, pretreated with
H2SO4 5% w/w, 175 C, 5 min and detoxified with Amorphoteca resinae ZN1 at
28 C, 4–7 days. Roukas and Kotzekidou (2020) obtained 351.5 g/kg dry pome-
granate peel by solid state fermentation, a promising method with low or no
pretreatment or complex equipment, although with a lower capacity to use nutrients
(Table 8.3).
8 Food Waste Biorefinery for Bioenergy and Value Added Products 207
Lactic acid (LA) is a very versatile compound that arouses much interest, accounting
its production through fermentative processes for around 80–90%. The global
demand for LA is estimated to increase from 1220 kt in 2016 to 1960 kt in 2025
(López-Gómez et al. 2019). The chemical route generates racemic mixtures of this
acid, requiring expensive separation processes. However, through microbial fermen-
tation, D- or L-lactic isomers can be generated. LA is widely used as acidulant in the
food industry. Furthermore, it is used in the manufacture of topical ointments,
lotions, parenteral solutions, cleaning agents, solvents, humectants and in the gen-
eration of oxygenated chemicals such as esters. In addition, polylactic acid (PLA) is
a green, biodegradable and biocompatible polymer obtained by ring opening from
lactide, a cyclic lactic acid dimer. Table 8.4 presents some LA production routes
from FW.
Among LA-producing microorganisms, the bacteria of the genus Lactobacillus
should be highlighted, being the most abundant genera. In recent years, numerous
studies have been carried out based on the use of FW as a cheap source of nutrients
(López-Gómez et al. 2019; Rawoof et al. 2020). Ahmad et al. (2020) studied the
effect of different pH and total solids content on dark fermentation of FWs without
inoculation of external microorganisms. The authors obtained a maximum LA
concentration of 12.87 g/L after having carried out an enzymatic pretreatment with
a cellulase at 50 C, pH 5–5.5 for 72 h. In order to reduce production costs, Jiang
et al. (2019) obtained 79.1 g/L of L-lactic acid through a simultaneous saccharifica-
tion and fermentation (SSF) in fed batch from 100 g/L of corncob residues
supplemented with 15 g/L peanut meal, 30 g/L CaCO3, 10 U/g cellulose, 15 U/g
glucosidase and 0.3 g/L protease. Through a continuous operation, Peinemann et al.
(2019) achieved a yield of 0.86 g/g with the addition of glucoamylase. However,
Costa et al. (2021) did not need to add any enzyme in their process, reaching a yield
208 M. Ladero et al.
of 78.3% in a batch operation from ricotta cheese whey (RCW) and centrifuged pear
processing (PP) residue in a 90:10 ratio.
Succinic acid (SA) is widely used in the food and pharmaceutical industry as an
additive, while it is an intermediate of the production of polybutylene succinate,
polyesters, polyols, resins, coatings and pigments in the traditional chemical indus-
try. In the bioeconomy era, SA is used mainly in the generation of intermediate
chemicals such as tetrahydrofuran, hydroxysuccinimide or maleic acid, among
others. In addition, the potential application of this acid in the manufacture of
biodegradable polymers should be highlighted (Dai et al. 2020). In 2017, the
bio-derived SA market was estimated at US$175.7 million and a compound annual
growth rate of 20% was forecasted, reaching a value of US$900 million by 2026
(Li et al. 2020). Table 8.5 features some studies on SA production starting from FW.
Bacteria isolated from the rumen of ruminants, including Actinobacillus.
succinogenes, Anaerobiospirillum succiniciproducens, and Mannheimia
succiniciproducens are the most promising candidates that can naturally produce
SA as the main product during anaerobic fermentation. However, some processes
have also been developed using fungi like Saccharomyces cerevisiae, Yarrowia
lipolytica, Byssochlamys nivea and Paecilomyces varioti. To achieve sustainable
production of bio-based SA, expensive sources of carbon and nitrogen, mainly
glucose and yeast extract, can be replaced by FW (Dai et al. 2020; Li et al. 2020).
For this reason, after optimizing the hydrolysis conditions, Li et al. (2018) used fruit
and vegetable waste (FVW) pretreated with 2% glucoamylase, 1% cellulase, 2%
hemicellulase and 0.25% pectinase at 55 C, pH 5, 300 rpm for 48 h. The authors
8 Food Waste Biorefinery for Bioenergy and Value Added Products 209
Fumaric acid (FA) and itaconic acid (IA) are included by the United States Depart-
ment of Energy on the list of top 12 building blocks to be obtained biotechnolog-
ically (Zhang et al. 2013). Both are known to be produced best using filamentous
fungi.
FA ((E)-2-butanodioic acid) production was linked to the Rhizopus genus until
50s in the XX century, when the petrochemical production route was developed.
Presently, FA is widely used on food and drug industries with promising applica-
tions in the polymer industry, having a global production of around 346 ktons/year
(Martin-Dominguez et al. 2018). This demand is satisfied through a petrochemical
process consisting of maleic acid isomerization. Very intensive conditions and the
presence of maleic anhydride are managed in this process, entailing a great environ-
mental risk. In an attempt to have fewer intensive requirements, enzymatic routes are
currently under study, although their maturity is still limited (Martin-Dominguez
210 M. Ladero et al.
et al. 2018). Table 8.6 summarizes some fumaric acid production processes using
FW as substrate.
In the development of a new fermentative process through the biorefinery con-
cept, several renewable raw materials related with food and agro-food industries
have been studied by several authors. In the last decade, many authors have tested
different waste types, achieving good production yields using “classical”
pre-treatments like drying and enzymatic hydrolysis, very consolidated on lignocel-
lulosic biomass revalorization (Deng and Aita 2018). Also, different kinds of
operation have been tested (Das et al. 2015b), such as Solid State Fermentation
(SSF), where excellent yields are reached having simpler pre-treatments and process
conditions. The direct use of FW has been studied, too, reaching very promising
results, mostly requiring easier handling pretreatments (Fan et al. 2020).
IA is employed overall as a very versatile monomer in the chemical industry,
particularly in fibres and polymers, also having many other applications in pharma
and health industries (Yang et al. 2020). It is known to be a much consolidated
bio-product obtained by a fermentative process carried out by Aspergillus terreus
using glucose or sucrose as the substrates (Huang et al. 2014b).
With a global production of about 407 kton/year, IA would involve a potential
market of US$567 million in 2022 (Ramakrishnan et al. 2020). Substrates are
responsible for its high production costs; thus, it is necessary to find an alternative
raw material to reduce prices (Huang et al. 2014b) and move this process into the
green industry cycle. Lignocellulosic biomass has been studied as potential raw
material for this process (Yang et al. 2020). As compiled in Table 8.7, very
promising production yields are reached using different FW.
8.3.6 Diols
Diols constitute a very versatile family of compounds for the chemical industry. This
term refers to compounds containing two alcohol groups that are attractive
biorefinery products mainly for their potential as monomers in bio-based polyesters
8 Food Waste Biorefinery for Bioenergy and Value Added Products 211
(Sato et al. 2020). Table 8.8 shows some examples of diols produced from FW
sources.
One of the most useful diols is 1,3-propanediol (1,3-PDO), whose main applica-
tion is being a monomer for polytrimethylene terephthalate production (Wang et al.
2019). 1,3-PDO fermentative production is usually through co-production with other
diols (Sato et al. 2020) or other bio-based molecules (Wang et al. 2019). Literature
on the production of 1,3-PDO from FW is scarce. However, the work by Wang
perfectly showcased the concept of biorefinery, using wasted cooking oil to obtain
biodiesel and raw glycerol as by-product, which can be ultimately used as substrate
for 1,3-PDO production (Wang et al. 2019).
2,3-butanediol (2,3-BDO) is a very promising diol with promising applications in
several processes such as polyurethane production (Psaki et al. 2019). Its
212 M. Ladero et al.
8.4.1 Biolubricants
Used oils like waste cooking oils are valuable sources of lipids that, under suitable
chemical transformations, can lead to obtain valuable chemicals. Specifically, lubri-
cants are an interesting family of compounds essential for tribological science, for
they are put to use to reduce overheating and friction in a variety of engines,
machinery, turbines and gear. In this sense, biolubricants pose an interesting oppor-
tunity due to their superior performance in terms of health and biodegradability
considerations. Vegetable oil-based bio-lubricants are recognized to have enhanced
lubricity, high viscosity, good anti-wear property, high viscosity index, increased
equipment service life, high load carrying ability, low evaporation rate, low emission
of metal traces into the atmosphere. To function as lubricants, waste oil must be
biologically or chemical processed (Esteban and Ladero 2018; Khodadadi et al.
2020). Table 8.9 presents some examples of the production of biolubricants from
FW.
Reactions of biolubricants manufacturing involve hydrolysis, transesterification,
epoxidation and oxirane ring-opening reactions (Li and Wang 2015; McNutt and He
2016). Both chemical and biological catalysis are applied. For example, the fatty
acid ethyl esters mixture, a fish oil residue obtained after the extraction of omega-3
polyunsaturated fatty esters, has been converted into mixtures of mono-, di-, and
triesters of trimethylolpropane by transesterification with a chemical catalyst
(Angulo et al. 2018). Many recent examples of the application of biocatalytic
reactions for the preparation of biolubricants can be found. In some cases trans-
formations involved enzymatic hydrolysis and chemically catalysed esterification to
obtain a maximum yield of biolubricant in minimum time (Chowdhury et al. 2013).
Oil can be produced from confectionery and wheat milling side streams. In that case,
nutrient-rich fermentation media were produced by a two-step bioprocess involving
crude enzyme production by solid state fermentation followed by enzymatic hydro-
lysis of confectionery industry waste. The extracted microbial oils were enzymati-
cally hydrolysed and the free fatty acids were esterified by Lipomod 34-MDP in a
solvent-free system with trimethylolpropane (TMP) and neopentyl glycol (NPG).
8 Food Waste Biorefinery for Bioenergy and Value Added Products 213
The highest conversion yields were 88% and 82.7% for NPG esters of R. toruloides
and C. curvatus, respectively (Papadaki et al. 2018).
Enzymatic transformation can be used for the synthesis of biolubricants from
by-products of soybean oil processing using the lipase from Candida rugosa (CRL)
both in free and immobilized forms, to produce biolubricants via their enzymatic
esterification with neopentyl glycol (NPG) and trimethylolpropane (TMP) alcohols
in a solvent-free medium (Fernandes et al. 2021). Another example is the production
of a bio-lubricant, which was obtained through esterification of WCOs free fatty
acids with neopentyl glycol by using Thermomyces lanuginosus lipase
(TL) immobilized on Fe3O4-CA (citric acid modified magnetite nanoparticles)
catalyst. The immobilized lipase was used for the biolubricant synthesis from
WCOs in a solvent-free system (Sarno et al. 2019). Immobilized enzyme catalysts
have also been used in esters from branched alcohols and dicarboxylic linear acids
(Serrano-Arnaldos et al. 2021). For example, a green and efficient strategy for the
preparation of octylated branched biolubricant from waste cooking oil involved
hydrolysis and esterification with 2-ethylhexanol employing lipase (Novozym
435) and further epoxidation and activation by a low-cost nucleophilic reagent,
octanoic acid, to prepare octylated branched biolubricant using an recyclable ionic
liquid, [HMIm][PF6], as catalyst (Zhang et al. 2020).
wide cellulose polymorphism I. Some bacteria genera are BC producers, among all
microorganisms Komagataeibacter sp. (known before as Gluconacetobacter) is the
most significant BC producer. Nevertheless, even though BC and PC have the same
molecular composition, they differ in their chemical and physical properties. BC has
high purity due to the absence of hemicellulose and lignin, high mechanical strength,
elevated water holding capacity, biocompatibility and high crystallinity (approx.
90%) make it an interesting material for a wide range of applications.
The most widespread culture medium for BC production is the Hestrin-Schramm
medium (HSM). The main drawbacks of HSM are the cost of carbon and nitrogen
sources and the requirement of a buffer to avoid the decrease of pH caused by the
gluconic acid release (Hussain et al. 2019). Depending on the mixing of culture
media, a pellicle in the broth surface can be obtained (most common) if it is statically
incubated or sphere-like particles, if submitted to stirring conditions. For BC pro-
duction, general conditions are 28–30 C, pH from 4.0 to 8.0 and an incubation
period from 3 to 25 days or more, in static conditions (Hussain et al. 2019).
Regarding the BC purification process, it involves an alkaline treatment followed
by a washing step. From an industrial point of view, BC generation is hindered by
both the considerable economic cost of the culture media (approx. 30% of process
economics) and productivity. Nowadays, there are companies that produce BC
employing coconut water as raw material and use batch reactors (tanks and trays)
for the production. Nonetheless, this source is not available worldwide. In the last
years, a variety of FW has been tested, reaching among 2.1–10.8 g/L from 3 to
28 days, depending on the raw material and culture condition (Fan et al. 2016;
Abdelraof et al. 2019; Kuo et al. 2019; Salari et al. 2019). Furthermore, pretreatment
steps are usually employed when FW is used in order to breakdown their polymers,
thereby enhancing carbon intake by the microorganisms. This approach shows
interesting results if the production of BC with HSM is compared to the higher
yield of BC achieved with FW (Table 8.10). The second possible step is to isolate or
modify BC producers using all the molecular biology, genetic and microbiology
tools available nowadays.
The broad diversity of BC application is caused by its remarkable physicochem-
ical properties added to its classification as GRAS (Generally recognized as safe) by
the FDA (Food and drug administration) in 1992. In the medical field, it has been
applied for the development of wound dress, dentary prostheses, tissue scaffolds,
blood vessel replacements and biosensors. In the food industry, it has been used as
packaging material (biopackaging), and a food ingredient like fat substituent,
probiotics immobilizer and stabilizing agent (Andriani et al. 2020). In the textile
industry, composite manufacturing, 3D printing material and optoelectronics are
examples of the extensive utilization list of this biopolymer (Andriani et al. 2020;
Hussain et al. 2019; Zhong 2020).
8 Food Waste Biorefinery for Bioenergy and Value Added Products 215
At present, PHAs comprise 1.7% of the 2.11 million tons of the global bioplastic
market. PHA production takes place using pure microbial cultures and definite
fermentation broth, which includes carbon sources such as pure sucrose and glucose,
organic acids (from 4 to 10 carbon atoms), alcohols and oils (Anjum et al. 2016).
With regards to production at pilot-scale and large-scale processes, to date fed-batch
production shows the highest productivities, despite these continuous and semi-
continuous reactors have already been tested (Sabapathy et al. 2020). Table 8.11
compiles a few examples of PHA production from FW.
Fermentation process parameters which are crucial in PHAs synthesis are the use
of elevated carbon to nitrogen ratio, control of pH and feedstock composition, all
together affect the production yield and polymer composition (Anjum et al. 2016).
However, the price of raw materials is one of the main disadvantages of current
PHAs production methodologies comprising up to 40–50% of the final production
cost. FW is an alternative substrate for cost-reduction in PHA production. For PHAs
synthesis, carbohydrate-rich or fatty acids-rich residues are adequate for producer
microorganisms. Moreover, pretreatments are often needed to enhance substrate
access and reduce the possible toxicity via detoxification steps. In that sense FW,
for instance, cheese whey (rich in lactose and proteins), cereal residues (e.g. wheat
straw and rice straw), sugar industry by-products (e.g. molasses), fat-rich residues
(e.g. frying oil and seeds with high oil level) are used to substitute expensive raw
materials. Final PHAs production yields obtained with this feedstock exceed 50%
8 Food Waste Biorefinery for Bioenergy and Value Added Products 217
PHA content per dry cell weight, and can reach more than 80% in some cases
(Table 8.11).
Another critical drawback in PHAs manufacturing is the accumulation inside the
microorganism as mentioned above, generating complex downstream processes
which involves cell lysis to extract PHAs. Purification costs can account around
30% of the process economy. These drawbacks result in a high price of PHAs around
5–6 EUR/kg, a disadvantage if compared to 0.8–1.5 EUR/kg for most common
petroleum-related plastics (El-Malek et al. 2020; Khatami et al. 2021; Sirohi et al.
2020).
References
Abdelraof M, Hasanin MS, El-Saied H (2019) Ecofriendly green conversion of potato peel wastes
to high productivity bacterial cellulose. Carbohydr Polym 211:75–83
Ahmad A, Banat F, Taher H (2020) Enhanced lactic acid production from food waste in dark
fermentation with indigenous microbiota. Biomass Conv Bioref. https://doi.org/10.1007/
s13399-020-00801-2
Alavijeh MK, Yaghmaei S, Mardanpour MM (2020) Assessment of global potential of biohydrogen
production from agricultural residues and its application in nitrogen fertilizer production.
Bioenergy Res 13(2):463–476
Algapani DE, Wang J, Qiao W, Su M, Goglio A, Wandera SM, Jiang M, Pan X, Adani F, Dong R
(2017) Improving methane production and anaerobic digestion stability of food waste by
extracting lipids and mixing it with sewage sludge. Biores Technol 244:996–1005
Allison BJ, Simmons CW (2017) Valorization of tomato pomace by sequential lycopene extraction
and anaerobic digestion. Biomass Bioenerg 105:331–341
Amato A, Becci A, Beolchini F (2020) Citric acid bioproduction: the technological innovation
change. Crit Rev Biotechnol 40(2):199–212
Ambat I, Srivastava V, Sillanpää M (2018) Recent advancement in biodiesel production method-
ologies using various feedstock: a review. Renew Sustain Energy Rev 90:356–369
Amiri H, Karimi K (2018) Pretreatment and hydrolysis of lignocellulosic wastes for butanol
production: challenges and perspectives. Bioresour Technol 270(December):702–721
Andriani D, Apriyana AY, Karina M (2020) The optimization of bacterial cellulose production and
its applications: a review. Cellulose 27:6747–6766
Angulo B, Fraile JM, Gil L, Herrerías CI (2018) Bio-lubricants production from fish oil residue by
transesterification with trimethylolpropane. J Clean Prod 202:81–87
Anjum A, Zuber M, Zia KM, Noreen A, Anjum MN, Tabasum S (2016) Microbial production of
polyhydroxyalkanoates (PHAs) and its copolymers: a review of recent advancements. Int J Biol
Macromol 89:161–174
Ayodele BV, Alsaffar MA, Mustapa SI (2020) Na overview of integration opportunities for
sustainable bioethanol production from first- and second-generation sugar-based feedstocks. J
Clean Prod 245:118857
Azwar MY, Hussain MA, Abdul-Wahab AK (2014) Development of biohydrogen production by
photobiological, fermentation and electrochemical processes: a review. Renew Sustain Energy
Rev 31:158–173
Babaki M, Yousefi M, Habibi Z, Mohammadi M (2017) Process optimization for biodiesel
production from waste cooking oil using multi-enzyme systems through response surface
methodology. Renew Energy 105:465–472
218 M. Ladero et al.
Degfie TA, Mamo TT, Mekonnen YS (2019) Optimized biodiesel production from waste cooking
oil (WCO) using calcium oxide (CaO) nano-catalyst. Sci Rep 9:18982
Deng F, Aita GM (2018) Fumaric acid production by Rhizopus oryzae ATCC® 20344™ from
lignocellulosic syrup. BioEnergy Res 11:330–340
Dhiman S, Mukherjee G (2020) Present scenario and future scope of food waste to biofuel
production. J Food Proc Eng e13594
El-Malek FA, Khairy H, Farag A, Omar S (2020) The sustainability of microbial bioplastics,
production and applications. Int J Biol Macromol 157:319–328
Esmaeili SAH, Sobhani A, Szmerekovsky J, Dybing A, Pourhashem G (2020) First-generation-
vs. second-generation: a market incentives analysis for bioethanol supply chains with carbon
policies. Appl Energy 277:115606
Esteban J, Ladero M (2018) Food waste as a source of value-added chemicals and materials: a
biorefinery perspective. Int J Food Sci Technol 53:1095–1108
Esteban J, Yustos P, Ladero M (2018) Catalytic processes from biomass-derived hexoses and
pentoses: a recent literature overview. Catalysts 8(12):637
Esteban J, Vorholt AJ, Leitner W (2020) An overview of the biphasic dehydration of sugars to
5-hydroxymethylfurfural and furfural: a rational selection of solvents using COSMO-RS and
selection guides. Green Chem 22(7):2097–2128
Fan X, Gao Y, He W, Hu H, Tian M, Wang K, Pan S (2016) Production of nano bacterial cellulose
from beverage industrial waste of citrus peel and pomace using Komagataeibacter xylinus.
Carbohydr Polym 151:1068–1072
Fan T, Liu X, Zhao R, Zhang Y, Liu H, Wang Z et al (2020) Hydrolysis of food waste by hot water
extraction and subsequent Rhizopus fermentation to fumaric acid. J Environ Manag 270:110954
FAOSTAT Home Page. http://www.fao.org/faostat/en/#data. Accessed 17 Jan 2021
Fernandes KV, Cavalcanti ED, Cipolatti EP, Aguieiras EC, Pinto MC, Tavares FA, da Silva PR,
Fernandez-Lafuente R, Arana-Pena S, Pinto JC, Assuncao CL (2021) Enzymatic synthesis of
biolubricants from by-product of soybean oil processing catalyzed by different biocatalysts of
Candida rugosa lipase. Catal Today 362:122–129
Ferone M, Ercole A, Raganati F, Olivieri G, Salatino P, Marzocchella A (2019) Efficient succinic
acid production from high-sugar-content beverages by Actinobacillus succinogenes. Biotechnol
Prog 35(5):e2863
Ferrero GO, Rojas HJ, Argaraña CE, Eimer GA (2016) Towards sustainable bio-fuel production:
design of a new biocatalyst to biodiesel synthesis from waste oil and commercial ethanol. J
Clean Prod 139:495–503
Fonseca JM, Teleken JG, de Cinque Almeida V, da Silva C (2019) Biodiesel from waste frying oils:
methods of production and purification. Energy Convers Manag 184:205–218
Food and Agriculture Organization of the United Nations (FAO). Goal 12: responsible consumption
and production home page. http://www.fao.org/sustainable-development-goals/indicators/1231/
en/. Accessed 15 Jan 2021
Foroutan R, Mohammadi R, Esmaeili H, Bektashi FM, Tamjidi S (2020) Transesterification of
waste edible oils to biodiesel using calcium oxide@magnesium oxide nanocatalyst. Waste
Manag 105:373–383
Franchetti M (2013) Economic and environmental analysis of four different con-figurations of
anaerobic digestion for food waste to energy conversion using LCA for: a food service provider
case study. J Environ Manag 123:42–48
Fu C, Li Z, Jia C, Zhang W, Zhang Y, Yi C, Xie S (2020) Recent advances on bio-based isobutanol
separation. Energy Conver Manag X:100059
Gebremariam SN, Marchetti JM (2018) Economics of biodiesel production: review. Energy
Convers Manag 168:74–84
Goh BH, Chong CT, Ge Y, Ong HC, Ng JH, Tian B, Ashokkumar V, Lim S, Seljak T, Józsa V
(2020) Progress in utilisation of waste cooking oil for sustainable biodiesel and biojet fuel
production. Energy Convers Manag 223:113296
220 M. Ladero et al.
Grand View Research (2016) GRV: isobutanol market size to reach $1.18 billion by 2022 [Online].
Available from https://www.grandviewresearch.com/press-release/global-isobutanol-market
Grigore ME, Grigorescu RM, Iancu L, Ion RM, Zaharia C, Andrei ER (2019) Methods of synthesis,
properties and biomedical applications of polyhydroxyalkanoates: a review. J Biomater Sci
Polym Ed 30(9):695–712
Gupta AR, Rathod VK (2018) Calcium diglyceroxide catalyzed biodiesel production from waste
cooking oil in the presence of microwave: optimization and kinetic studies. Renew Energy 121:
757–767
Gupta J, Agarwal M, Dalai AK (2020) An overview on the recent advancements of sustainable
heterogeneous catalysts and prominent continuous reactor for bio-diesel production. J Ind Eng
Chem 88:58–77
Hajjari M, Tabatabaei M, Aghbashlo M, Ghanavati H (2017) A review on the prospects of
sustainable biodiesel production: a global scenario with an emphasis on waste-oil biodiesel
utilization. Renew Sustain Energy Rev 72:445–464
Hamza M, Ayoub M, Shamsuddin RB, Mukhtar A, Saqib S, Zahid I, Ameen M, Ullah S, Al-Sehemi
AG, Ibrahim M (2020) A review on the waste bio-mass derived catalysts for biodiesel produc-
tion. Environ Technol Innov 101200
Han W, Yan Y, Gu J, Shi Y, Tang J, Li Y (2016) Techno-economic analysis of a novel bioprocess
combining solid state fermentation and dark fermentation for H2 production from food waste.
Int J Hydrog Energy 41(48):22619–22625
Hitam CNC, Jalil AA (2020) A review on biohydrogen production through photo-fermentation of
lignocellulosic biomass. Biomass Convers Biorefinery 1–19
Hou W, Bao J (2018) Simultaneous saccharification and aerobic fermentation of high titer cellulosic
citric acid by filamentous fungus Aspergillus niger. Bioresour Technol 253:72–78
Huang X, Chen M, Lu X, Li Y, Li X, Li J (2014a) Direct production of itaconic acid from liquefied
corn starch by genetically engineered Aspergillus terreus. Microb Cell Fact 13:108
Huang X, Lu X, Li Y, Li X, Li JJ (2014b) Improving itaconic acid production through genetic
engineering of an industrial Aspergillus terreus strain. Microb Cell Factories 13(1):1–9
Hussain Z, Sajjad W, Khan T, Wahid F (2019) Production of bacterial cellulose from industrial
wastes: a review. Cellulose 26:2895–2911
Huzir NM, Aziz MMA, Ismail SB, Abdullah B, Mahmood NAN, Umor NA, Muhammad SAFAS
(2018) Agro-industrial waste to biobutanol production: eco-friendly biofuels for next genera-
tion. Renew Sustain Energy Rev 94:476–485
Ilhak, MI, Tangoz S, Akansu SO, Kahraman N (2019) Alternative fuels for internal combustion
engines in the future of internal combustion engines. IntechOpen. https://doi.org/10.5772/
intechopen.85446
International Energy Agency Home Page. https://www.iea.org/data-and-statistics?
country¼WORLD&fuel¼Energy%20supply&indicator¼TPESbySource. Accessed
17 Jan 2021
International Energy Association. Outlook for biogas and biomethane. Prospects for organic
growth. World Energy Outlook Special Report. Accessed in January 2021 at https://www.iea.
org/reports/outlook-for-biogas-and-biomethane-prospects-for-organic-growth/an-introduction-
to-biogas-and-biomethane
Jeevan Kumar SP, Sampath Kumar NS, Chitagunta AD (2020) Bioethanol production from cereal
crops and lignocellulosic-rich agro-residues: prospects and challenges. SN Appl Sci 2:1673
Ji Q, Yu X, Chen L, Yagoub AEA, Zhou C (2020) Aqueous choline chloride/γ-valerolactone as
ternary green solvent enhance Al(III)-catalyzed hydroxymethylfurfural production from rice
waste. Energy Technol 8(12):2000597
Jiang S, Xu P, Tao F (2019) L-Lactic acid production by Bacillus coagulans through simultaneous
saccharification and fermentation of lignocellulosic corncob residue. Bioresour Technol Rep 6:
131–137
Jin C, Yao M, Liu H, Chia-fon FL, Ji J (2011) Progress in the production and application of
n-butanol as a biofuel. Renew Sustain Energy Rev 15(8):4080–4106
8 Food Waste Biorefinery for Bioenergy and Value Added Products 221
Karimi S, Karimi K (2018) Efficient ethanol production from kitchen and garden wastes and biogas
from the residues. J. Clean Prod 187:37–45
Khalid A, Aslam M, Qyyum MA, Faisal A, Khan AL, Ahmed F, Lee M, Kim J, Jang N, Chang IS,
Bazmi AA (2019) Membrane separation processes for dehydration of bioethanol from fermen-
tation broths: recent developments, challenges, and prospects. Renew Sustain Energy Rev 105:
427–443
Khatami K, Perez-Zabaleta M, Owusu-Agyeman I, Cetecioglu Z (2021) Waste to bioplastics: how
close are we to sustainable polyhydroxyalkanoates production? Waste Manag 119:374–388
Khodadadi MR, Malpartida I, Tsang C-W, Lin CSK, Len C (2020) Recent advances on the catalytic
conversion of waste cooking oil. Mol Catal 494:111128
Kim S, Lee Y, Lin KYA, Hong E, Kwon EE, Lee J (2020) The valorization of food waste via
pyrolysis. J Clean Prod 259:120816
Kirubakaran M, Arul Mozhi Selvan V (2018) A comprehensive review of low cost biodiesel
production from waste chicken fat. Renew Sustain Energy Rev 82:390–340
Kourmentza C, Costa J, Azevedo Z, Servin C, Grandfils C, De Freitas V, Reis MAM (2018)
Burkholderia thailandensis as microbial cell factory for the bioconversion of used cooking oil to
polyhydroxyalkanoates and rhamnolipids. Bioresour Technol 247:829–837
Kovalcik A, Pernicova I, Obruca S, Szotkowski M, Enev V, Kalina M, Marova I (2020) Grape
winery waste as a promising feedstock for the production of polyhydroxyalkanoates and other
value-added products. Food Bioprod Process 124:1–10
Kulkarni NV, Brennessel WW, Jones WD (2018) Catalytic upgrading of ethanol to n-butanol via
manganese-mediated Guerbet reaction. ACS Catal 8(2):997–1002
Kuo CH, Huang CY, Shieh CJ, Wang HMD, Tseng CY (2019) Hydrolysis of orange peel with
cellulase and pectinase to produce bacterial cellulose using Gluconacetobacter xylinus. Waste
Biomass Valor 10:85–93
Lam CM, Iris KM, Hsu SC, Tsang DC (2018) Life-cycle assessment on food waste valorisation to
value-added products. J Clean Prod 199:840–848
Lawrence Livermore National Laboratory. LLNL Flowcharts Home Page. https://flowcharts.llnl.
gov/. Accessed 17 Jan 2021
Li W, Wang X (2015) Bio-lubricants derived from waste cooking oil with improved oxidation
stability and low-temperature properties. J Oleo Sci 64(4):367–374
Li C, Yang X, Gao S, Chuh AH, Lin CSK (2018) Hydrolysis of fruit and vegetable waste for
efficient succinic acid production with engineered Yarrowia lipolytica. J Clean Prod 179:151–
159
Li C, Ong KL, Yang X, Lin CSK (2019) Bio-refinery of waste streams for green and efficient
succinic acid production by engineered Yarrowia lipolytica without pH control. Chem Eng J
371:804–812
Li C, Ong KL, Cui Z, Sang Z, Li X, Patria RD, Qi Q, Fickers P, Yan J, Lin CSK (2020) Promising
advancement in fermentative succinic acid production by yeast hosts. J Hazard Mater 401:
123414
Liu H, Ma J, Wang M, Wang W, Deng L, Nie K, Yue X, Wang F, Tan T (2016) Food waste
fermentation to fumaric acid by Rhizopus arrhizus RH7-13. Appl Biochem Biotechnol 180:
1524–1533
Liu X, Lv J, Xu J, Xia J, He A, Zhang T, Li X, Xu J (2018) Effects of osmotic pressure and pH on
citric acid and erythritol production from waste cooking oil by Yarrowia lipolytica. Eng Life Sci
18(6):344–352
Loizidou M, Alamanou DG, Sotiropoulos A, Lytras C, Mamma D, Malamis D, Kekos D (2017)
Pilot scale system of two horizontal rotating bioreactors for bioethanol production from
household food waste at high solid concentrations. Waste Biomass Valor 8(5):1709–1719
López-Gómez JP, Alexandri M, Schneider R, Venus J (2019) A review on the current developments
in continuous lactic acid fermentations and case studies utilising inexpensive raw materials.
Process Biochem 79:1–10
222 M. Ladero et al.
Lucas-Torres C, Lorente A, Cabañas B, Moreno A (2016) Microwave heating for the catalytic
conversion of melon rind waste into biofuel precursors. J Clean Prod 138:59–69
Lui MY, Wong CYY, Choi AWT, Mui YF, Qi L, Horváth I (2019) Valorization of carbohydrates of
agricultural residues and food wastes: a key strategy for carbon conservation. ACS Sustain
Chem Eng 7(21):17799–17807
Martin-Dominguez V, Estevez J, Ojembarrena FB, Santos VE, Ladero M (2018) Fumaric acid
production: a biorefinery perspective. Fermentation 4:33
McNutt J, He Q (2016) Development of biolubricants from vegetable oils via chemical modifica-
tion. J Ind Eng Chem 36:1–12
Merza F, Fawzy A, AlNashef I, Al-Zuhair S, Taher H (2018) Effectiveness of using deep eutectic
solvents as an alternative to conventional solvents in enzymatic biodiesel production from waste
oils. Energy Rep 4:77–83
Mohan SV, Dahiya S, Amulya K, Katakojwala R, Vanitha TK (2019) Can circular bioeconomy be
fueled by waste biorefineries—A closer look. Bioresour Technol Rep 7:100277
Mores S, de Souza Vandenberghe LP, Júnior AIM, de Carvalho JC, de Mello AFM, Pandey A,
Soccol CR (2021) Citric acid bioproduction and downstream processing: status, opportunities,
and challenges. Bioresour Technol 124426
Ntaikou I, Antonopoulou G, Lyberatos G (2010) Biohydrogen production from biomass and wastes
via dark fermentation: a review. Waste Biomass Valorization 1(1):21–39
OECD/FAO (2020) Agricultural Outlook (edition 2020) data base. https://stats.oecd.org/
BrandedView.aspx?oecd_bv_id¼agr-data-en&doi¼4919645f-en
OECD/Food and Agriculture Organization of the United Nations (2018) Chapter 9. Biofuels in
OECD-FAO Agricultural Outlook 2018-2027, Paris/Food and Agriculture Organization of the
United Nations, Rome. https://doi.org/10.1787/agr_outlook-2018-12-en
OHair J, Jin Q, Yu D, Wu J, Wang H, Zhou S, Huang H (2021) Non-sterile fermentation of food
waste using thermophilic and alkaliphilic Bacillus licheniformis YNP5-TSU for 2,3-butanediol
production. Waste Manag 120:248–256
Ong KL, Li C, Li X, Zhang Y, Xu J, Lin CSK (2019) Co-fermentation of glucose and xylose from
sugarcane bagasse into succinic acid by Yarrowia lipolytica. Biochem Eng J 148:108–115
Ozdal M, Kurbanoglu EB (2019) Citric acid production by Aspergillus niger from agro-industrial
by-products: molasses and chicken feather peptone. Waste Biomass Valor 10(3):631–640
Papadaki A, Fernandes KV, Chatzifragkou A, Aguieiras ECG, da Silva JAC, Fernandez-Lafuente R
et al (2018) Bioprocess development for biolubricant production using microbial oil derived via
fermentation from confectionery industry wastes. Bioresour Technol 267:311–318
Paudel S, Kang Y, Yoo YS, Seo GT (2017) Effect of volumetric organic loading rate (OLR) on H2
and CH4 production by two-stage anaerobic co-digestion of food waste and brown water. Waste
Manag 61:484–493
Peinemann JC, Demichelis F, Fiore S, Pleissner D (2019) Techno-economic assessment of
non-sterile batch and continuous production of lactic acid from food waste. Bioresour Technol
289:121631
Pereira JR, Araújo D, Freitas P, Marques AC, Alves VD, Sevrin C, Grand-fils C, Fortunato E, Reis
MAM, Freitas F (2021) Production of medium-chain-length polyhydroxyalkanoates by Pseu-
domonas chlororaphis subsp. aurantiaca: cultivation on fruit pulp waste and polymer charac-
terization. Int J Biol Macromol 167:85–92
Poppe JK, Matte CR, Fernandez-Lafuente R, Rodrigues RC, Ayub MAZ (2018) Transesterification
of waste frying oil and soybean oil by combi-lipases under ultrasound-assisted reactions. Appl
Biochem Biotechnol 186(3):576–589
Psaki O, Maina S, Vlysidis A, Papanikolaou S, Machado de Castro A, Freire DMG, Dheskali E,
Kookos I, Koutinasa A (2019) Optimisation of 2,3-butanediol production by Enterobacter
ludwigii using sugarcane molasses. Biochem Eng J 152:107370
Pugazhendhi A, Mathimani T, Varjani S, Rene ER, Kumar G, Kim SH, Ponnusamy VK, Yoon JJ
(2019) Biobutanol as a promising liquid fuel for the future-recent updates and perspectives. Fuel
253:637–646
8 Food Waste Biorefinery for Bioenergy and Value Added Products 223
F. Guilayn (*)
SUEZ, CIRSEE, Le Pecq, France
e-mail: [email protected]
J. Jimenez
LBE, Univ. Montpellier, INRAE, Narbonne, France
F. Monlau
APESA, Pôle Valorisation, Cap Ecologia, Lescar, France
C. Vaneeckhaute
BioEngine, Chemical Engineering Department, Université Laval, Québec, QC, Canada
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 227
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_9
228 F. Guilayn et al.
Abbreviations
9.1 Introduction
Improving soil quality and fertility by recycling organic waste products has become
a crucial approach to substitute mineral chemicals and to participate in climate
change mitigation by promoting carbon storage in soils. (Alburquerque et al.
2012; Gissén et al. 2014; Minasny et al. 2017). In this context anaerobic digestion
(AD) should play a crucial through the valorisation of digestate, the digested output.
This process is being successfully implemented in Europe as shown by the 18,202
plants recorded in 2018 by the European Biogas Association (EBA 2019).
AD consists of several biological reactions where the organic carbon is converted
to its most oxidised state (CO2) and its most reduced form (CH4). The methane is an
energy vector for directly producing electricity, heat and power or after purification
and injection in the natural gas grid. In an energetic and climatic crisis context, AD is
considered as an alternative for organic waste valorisation.
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 229
The use and development of digestate upgrading technologies is crucial and have
to take into account several objectives including (1) removing undesired digestate
characteristics (e.g. through chemical precipitation, nitrification/denitrification,
impurity removal), (2) reaching disposal requirements, (3) reducing logistics costs
by concentrating the valuable components (e.g. liquid/solid separation technologies
more or less advanced, drying, evapo-concentration); (4) producing high market-
value products (e.g. N-stripping, struvite crystallisation, biostimulants, etc.)
(5) ensuring homogenous quality over time; and (6) increasing market acceptance
and (7) creating new markets (Guilayn et al. 2020b).
In this chapter, an overview of digestates valorisation pathways is provided from
full-scale to the most promising approaches under development. First, the effects of
AD and the direct agronomical valorisation of digestate are discussed. Then, the
most relevant value-added products that can be obtained from digestates, according
to the author’s view, are presented and discussed. Finally, insights on the regulatory
framework in the EU and in North America are presented.
As AD will produce a series of systematic effects on its feedstock, several trends can
be generalized regarding the quality of digestates. A simplified overview of AD effet
on input is provided in Fig. 9.1.
Fig. 9.1 A simplified summary of most notable effects of AD. Areas are proportional to mass and
illustrate the case of a feedstock containing about 30% total solids
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 231
Due majorly to the conversion of OM to biogas and to a lesser extent due to the
consumption of water for hydrolysis, digestate mass and volume are always reduced
compared to the AD reactor input. As most of this mass loss is the biodegradation of
organic particles, the greater the dry matter (DM) content of the input, the more this
effect will be noticeable. For instance, full-scale Dry-AD can result in up to 10–25%
mass reduction (as 1-input/output) (Bolzonella et al. 2006; Schievano et al. 2011;
Banks et al. 2011). However, it is important to consider the overall process line,
notably considering AD pre-treatment. As an example, in the case of biowaste
wet-AD, the process line might rely on heavy dilution for depackaging and/or
hygienisation steps which can increase by two- to threefold the wet volume of AD
input compared to the plant feedstock (data from field expertise).
9.2.4 Innocuity
In the EU, agricultural digesters are the most common, notably in Germany, due to
historical support policies. According to the European Biogas Association, there
were nearly 18,000 biogas plants in Europe in 2017 (landfills included), about 50%
of them being agricultural digesters in Germany. Depending on member state’s
national policy, typical inputs may vary. For instance, in France, agricultural AD
will typically co-digest manure/slurry, crop residues and/or catch crops. In Germany,
the legislation historically allowed AD of energy crops as the main crop cultivated
specifically for AD. These digesters are usually mesophilic continuous stirred-tank
reactors (CSTR) but agricultural dry-AD is also reported, especially for fibrous
manure. In Fig. 9.2, they can be related to digestate types 1 and 6. Higher dry matter
(>5% in Wet-AD, relating to interquartile ranges), volatile solids (>70% DM) and
C/N ratio (5–20) would be related to the more fibrous characteristic of agricultural
inputs. Agricultural digestates tend to be easier to commercialize from a regulatory
point-of-view. This is the most dominant digestate type as, in Europe, most biogas
installations are based on dedicated energy crops (Brémond et al. 2021). This trend
tends to continue but probably shift to rotational crops (catch crops) to avoid food
production competition while providing side benefits to the main crop (Brémond
et al. 2021).
Wastewater treatment plant (WWTP) digesters are also numerous. Given the low
concentration of sewage sludge, they are usually large mesophilic CSTR. As the
sludge is conventionally digested prior to dewatering, these AD plants are usually on
234 F. Guilayn et al.
Fig. 9.2 Digestate typology based on a clustering analysis on a database of 91 samples (extracted
from Guilayn et al. 2019a)
The organic fraction of municipal solid waste (OFMSW) is usually treated in large-
scale municipal biogas facilities. OFMSW is obtained after a mechanical biological
treatment (MBT) of mixed-collection urban waste. MBT of such a stream usually
produces a fibrous cellulosic biodegradable material with a structured bulk capacity
without needing a structuring co-substrate. MBT is thus often followed by a Dry-AD
thermophilic digester.
Another emerging configuration is Dry-AD of biowaste (source-separated) when
it contains a significant contribution of green waste serving as structuring material. It
is important to mention that some authors will also refer to source-separated
biowaste from municipal solid waste collection as OFMSW. In Fig. 9.2, they
compose most of type 3 and 5 digestates presenting low nutrient content that can
be associated to the typical cellulosic material obtained from MBT. In Europe, on the
one hand, the restrictions and end-of-waste scenarios described for sludge digestate
apply similarly to MBT digestate from mixed-collection waste. On the other hand,
source-separated biowaste digestate is clearly trending to reach more easily “end-of-
waste” status. For instance, biowaste (source-separated) was included in the latest
EU regulation on fertilizers (CE 2019/1009)
To conclude, even if AD configurations can be somehow generalized, the pro-
cess/digestate typology is also a result from a limited technical expertise imposing
design and operational rules that are being constantly challenged and optimized by
research. This process might take a significant time to reach the market, thus the
typology can be expected to be constantly evolving.
As organic matter is the major digestate component (other than moisture), valorizing
digestate organic matter is one of the greatest levers for creating value. In the future,
a more defined economic value to stabilized carbon is expected through carbon
credits. Additionally, an increasing economic value tends to be associated for the
beneficial effect of organic matter in agriculture, especially as most soils are being
rapidly depleted in OM. In this scenario, optimizing AD for considering not only
biogas yield but digestate OM quality and quantity can be crucial. A change of
paradigm will be needed. Adapting the process line to enhance final digestate quality
while lowering biogas production has already been done to achieve better economic
results in one of the largest urban AD facilities in Europe (internal information
from SUEZ).
First of all, feedstock quality plays a major role in the proportion of OM degraded
during the anaerobic digestion process and during land spreading of the digestate
(Guilayn et al. 2019a, 2020b). As previously said, fibers-like feedstock enhance OM
236 F. Guilayn et al.
in digestate because lignin-like compounds are not biodegraded (i.e., manure, crop
residues, straw), whereas biowaste-like feedstock decrease OM in digestate because
the OM content is more biodegradable (i.e., OFMSW and biowastes).
All bibliographic sources agree that a fraction of the digestate’s OM remains
biodegradable once it is incorporated into the soil, making post-treatment interesting
by composting to achieve a more stabilized OM. This fraction of biodegradable OM
converted into methane and carbon dioxide is dependent on the efficiency of AD
performance in particular on parameters such as temperature, organic loading rate
and hydraulic residence time. On average, 2/3 of the biodegradable OM introduced
into the digester is transformed into biogas (Alburquerque et al. 2012).
In addition to feedstock and process parameters, the technology itself seems to
have an impact on the soil amending value of digestates: those resulting from
dry-AD seem to contain more OM consisting mainly of carbon chains very difficult
to degrade by the microorganisms. AD would not decompose the humic potential of
OM because the microorganisms do not break down lignin, complex lipids, steroids
and precursor molecules for humus formation (Jimenez et al. 2017). As previously
stated, some authors indicate the presence of humification-like processes occurring
in AD (Brunetti et al. 2012), but humification is a controversial topic for over a
century in soil science. However, a stabilized OM is obtained in the digestate due to
the transformation of labile fractions of OM in biogas and the formation of more
stable compounds (increased degree of aromaticity, accumulation of longer aliphatic
chains) according to Möller (2015).
To conclude, the digestates contain a fraction of residual biodegradable OM and a
proportion of stabilized pre-humic matter, which depends on the conditions of the
digestion and feedstocks but also on the post-treatment applied to the digestate.
These parameters can be used as actuators to control the quality of the digestate.
The most classical valorisation of digestates and by-products is the agronomic use in
crop soils. Indeed, as previously mentioned, because of the stabilized OM and the
mineralized nutrients, digestates are considered as organic fertilizers if they are
properly managed. Three types of fertilisation can be obtained using organic resi-
dues in crop soils: (1) chemical fertilizing (available nutrients NPK used at short/
mid- term to feed the plants), (2) biological fertilizing (labile OM able to feed the soil
microorganisms at short term) and (3) physical fertilizing (amendment potential of
stabilized OM to improve the soil structure). Several different types of digestates
have been demonstrated to be effective replacements for mineral fertilizers while
reducing environmental impacts (Walsh et al. 2012; Koszel and Lorencowicz 2015;
Panuccio et al. 2019; Verdi et al. 2019). Nevertheless, regulatory constraints, good
management and spreading practices are necessary for controlling environmental
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 237
and health risks as it can be expected for any chemical fertilizers, raw manure and
sludge (Nkoa 2014).
Considering chemical fertilisation, mineral elements increase chemical fertility
by feeding the plants’ nutritional needs. Nitrogen fertilizing is the main target when
chemical fertilizer is considered. During AD, ammonia reacts with water in the
anaerobic environment to form ammonium found in the digestate. After land spread-
ing, nitrification by soil bacteria transforms ammonium into nitrate, which can be
assimilated by the plant, within a period of a few days to a few weeks. Losses in the
form of nitrous oxide or nitrogen oxide may occur during this process. However, the
total nitrogen stock remains constant compared to the undigested inputs, but the final
concentrations are higher, due to the loss of a quantity of carbon in the form of
biogas during the AD process (Tambone et al. 2009).
When the digestate is incorporated to the soil, part of the carbon is used as an
energy source for the soil microflora and nitrogen is used as a nutrient. A phenom-
enon of reorganization of mineral nitrogen is often observed: as soon as the digestate
is added to the soil, the microbial activity of the soil is reinforced with the production
of new cells because its incorporation provides a source of carbon energy. This
requires a proportional amount of nitrogen, which is part of the composition of many
essential molecules in cells. If the supplied OM does not contain enough of N to meet
this demand, microorganisms will take (and therefore immobilize) N soil solution to
be able to grow (Reibel and Leclerc 2018).
For raw and liquid digestate, with a lower C to N ratio compared to solid
digestate, studies report that the rate of net nitrogen mineralisation is usually positive
(Cavalli et al. 2016; Jimenez et al. 2020a). On the other hand, for the solid digestate,
net immobilisation of nitrogen, even in the medium term can occur, depending on
the digestate typology (Jimenez et al. 2020a). Indeed, the authors showed that the C
to N ratio is negatively correlated with soil mineralized N and N recovered in plant
tissues. It seems to be a good indicator for the mineralisation potential of organic N
and for apparent recovery fraction of N by plants as shown by Jimenez et al. (2020a).
As previously discussed, the nitrogen fertilizing value of digestate depends on:
(1) the mineral nitrogen content, which depends on the feedstocks and post-
treatments. More than half of the organic nitrogen contained in the raw digested
inputs is found as ammonium quickly mobilized by plants, conferring a strong
fertilizing potential. (2) The rate of mineralisation of the organic fraction of nitrogen
in digestates. The mineralisation kinetics depends on the type of feedstock and the
environmental conditions (temperatures and mainly soil moisture). However, the
mineralisation of organic nitrogen appears to play a minor role in the fertilizing
value. Little organic nitrogen remains in the soil after a growing season, thus
reducing the long-term effect of so-called residual nitrogen as well as long-term
nitrogen leaching (Reibel and Leclerc 2018). (3) The phenomenon of ammonia
volatilisation, which can in some cases reach 100% mineral nitrogen. This phenom-
enon is also a function of the type of feedstock, post-treatment, storage, spreading
and pedoclimatic conditions (Reibel and Leclerc 2018). The spreading conditions
have a strong influence on the apparent recovery fraction of N by crops: a liquid
digestate applied in surface can lose 35–100% of its ammonia (Nyord et al. 2008;
238 F. Guilayn et al.
AD enhances undoubtedly the N availability for plants growth, but this process can
also enhance the potential loss of ammonia by volatilisation during storage (mini-
mized if the storage is covered) and spreading compared to initial materials (Reibel
2018). This increase can be partially offset by the better infiltration capacity of liquid
digestates in the soil and by covering actions during storage as well as appropriate
practices during spreading such as rapid burial, dry climatic conditions and cool
temperatures, non-bare soil, etc. Spreading recommendations are found more suit-
able for slurries and is not observed for solids, composted and dried digestates.
Greenhouse gases can also be produced during storage and spreading. However,
some studies showed that AD decrease CH4 and N2O emissions compared with
feedstock spreading (Holly et al. 2017; Reibel 2018).
Another limitation of some types of digestates is the metallic trace elements
content. AD tends to increase trace mineral concentrations via degradation of the
organic matrix with a clearer tendency for sewage sludge and OFMSW. For the
digested sewage sludge, speciation of metallic trace elements seems to be evolving
towards more stable forms and less available during digestion (Reibel 2018).
Concerning organic micropollutants, AD modifies the concentrations of organic
contaminants. For example, the anaerobic digestion of urban sludge has demon-
strated a potential for degradation of PAH, nonylphenol and PCB (Reibel 2018).
According to Guilayn et al. (2020b), digestate farmland spreading is the mainly
digestate management strategy. However, for the non-agricultural digestate pro-
viders, several legal and technical bottlenecks are found as follows: (1) digestate
240 F. Guilayn et al.
whereas the LF is better pumpable. The SF might contribute to the soil bulking
capacity, whereas the LF has better infiltration in soil. Regarding post-treatment
(later detailed), the SF can be post-treated by technologies needing a low-moisture
content (notably composting and drying) whereas the LF through technologies
needing low-solids and/or presenting clogging issues (notably nitrogen stripping,
struvite recovery, (vacuum-)evaporation and membrane filtration).
Adding to this “physical” effect, digestate phase separation enhances nutrient and
organic matter management in agriculture. The SF concentrates 30–90% of the OM
into 5–50% of the initial mass (Guilayn et al. 2019b). It can thus be better used as a
soil amendment (adding to the bulking capacity effect), composted or as constituting
growth medium material replacing peat (Nesse et al. 2019). The LF concentrates
most of the ammoniacal nitrogen and potassium, while phosphorus tends to follow
the organic matter and tends thus to be concentrated in the SF (Guilayn et al.
2019a, b). This nutrient fractioning effect can be used as a strategy for improving
nutrient management (e.g. different crops or seasons) and increasing the spreading
surface of the fractions by reducing the concentration of the limiting nutrient (usually
N or P).
Given the benefits described above, digestate post-treatment technologies are clas-
sified in this chapter by their applicability concerning digestate rheological state. To
the best of the author’s knowledge, this process classification has not yet been linked
to an actual rheological objective parameter, but the DM content (i.e., total solids) is
a usual practical indicator. The techniques described below for liquid digestates are
usually applied to LF or wet-AD digestates with less than 5–8% DM. After phase
separation, adding to the enhanced physical aspects described in the previous
section, it is important to mention that most of the “liquid-state” technologies
described below will benefit from an enhanced nutrient content (e.g., N-stripping,
microalgae, membrane filtration and evaporation).
full-scale system called the Biorek Process has been installed by BIOSCAN (Den-
mark). Besides reversed osmosis, microfiltration and ultrafiltration units can also be
combined with ammonia stripping (Vaneeckhaute et al. 2017). Operating tempera-
tures for membrane filtration typically range between 10 and 40 C. The pH is
typically between 6 and 8. The cost of membrane filtration for manure and digestate
treatment was evaluated through a large-scale pilot project conducted at eight
installations in the Netherlands. The estimated cost ranged between 9 and 13 €/t of
manure or digestate (de Hoop et al. 2011). Potential revenues for the reversed
osmosis concentrates were estimated at an average of 6.1 €/t. An important technical
issue when using membrane filtration is clogging and fouling of the membrane,
which can result in significant costs for the use of cleaning products. In order to
mitigate this problem, Vaneeckhaute et al. (2012, 2019) evaluated the performance
of vibratory membrane filtration (VSEP technology) for digestate purification at
demonstration scale. Although the chemical use was significantly reduced, the
energy requirements of this technology are not negligible. The potential pollution
of the recovered concentrate flows, for example with heavy metals, also remains an
important point of attention.
Several technologies have been developed in the last decades to extract pure and
formulated end products from liquid fractions of digestate. Based on a review
executed by Vaneeckhaute et al. (2017), the current most implemented technologies
at full-scale include struvite precipitation and crystallisation, as well as ammonia
stripping and scrubbing.
Struvite precipitation allows to recover soluble phosphorus (ortho-P) from waste-
water streams at high phosphorus loads (usually >80 kg P/day) (Vaneeckhaute et al.
2018). It typically requires the addition of a magnesium source (typically MgO, Mg
(OH)2 or MgCl2) and caustic soda (NaOH) to increase the pH (8.3–10) and induce
precipitation of MgNH4PO46H2O (struvite). Alternatively, the pH can be increased
through stripping of carbon dioxide, without the addition of NaOH. The
precipitation-crystallisation reaction usually takes place in a fluidized bed reactor
or a completely stirred tank reactor. Important parameters to control include pH,
temperature, mixing intensity, and the presence of impurities such as calcium.
Struvite can be reused as a valuable slow-release fertilizer product containing the
macronutrients N, P and Mg that support plant growth, equivalent to 12% P (28%
P2O5). Capital costs for producing struvite have been estimated at 0.05 €/day.
However, the overall benefits for the treatment plant can reach 0.52 €/day
(Vaneeckhaute et al. 2017). The latter is related to chemical savings (conventional
Fe/Al-salts), savings from reduced sludge handling and disposal, as well as savings
related to the reduced cleaning of uncontrolled struvite deposits. The market value
for struvite is variable, with numbers of 45 €/t in Belgium to 250 €/t in Japan
(Vaneeckhaute et al. 2017).
Nitrogen stripping and scrubbing allows to recover nitrogen from a variety of
wastewater streams at high ammonia (NH3) nitrogen concentrations (usually
>500 mg/L) (Vaneeckhaute et al. 2018). It involves the physical transfer of ammo-
nia from the aqueous phase to the gas phase at increased pH. The ammonia is then
absorbed in sulphuric acid by means of an air scrubber to produce liquid ammonium
sulphate. Ammonia stripping typically takes place in a packed bed tower or a
submerged aeration system. Key advantages of the first is its easy operation and
low capital cost, while key advantages of the latter are the absence of packing
material (no risk of fouling) and its ability to increase the pH without chemical
addition (through CO2 stripping). Acidic air scrubbing usually takes place in a
packed bed tower. Sulphuric acid and air are introduced into the tower in counter-
current. Important parameters to control are pH (typically 10–11) and temperature
(typically 60–70 C). The ammonium sulphate concentration in the recovered liquid
product ranges from 25 to 40% by weight. It can be valorised as a fertilizer product
rich in the macronutrients nitrogen and sulphur. Capital costs for stripping are
estimated between 500,000 euros and 1.58 million euros for a packed bed tower,
whereas this amounts to 3.5–15 M€ for a submerged aeration system. The opera-
tional costs range between 1.4 and 2.5 M€/year (Vaneeckhaute et al. 2017). The
market value of the recovered ammonium sulphate solution ranges between 80 and
244 F. Guilayn et al.
120 €/t of fresh weight. As such, under optimal process conditions, nitrogen strip-
ping and scrubbing can be viable.
2015; Ronga et al. 2019a). Microalgae can also be used for the upgrading of biogas
and several studies have been done in the last decades (Tongprawhan et al. 2014;
Rodero Raya et al. 2019). Finally, when considering the valuation chains it is
important to consider the contamination risks (i.e. pathogens, organic and inorganic
contaminants) and biomass safety concerns (Markou et al. 2018) and for this purpose
human-consumption applications are often not considered. Up to date, some R&D
projects have been implemented in Europe coupling liquid digestate with microalgae
cultivation. Among these projects, the AlgaeBiogas project (https://algaebiogas.eu/)
has investigated the use of liquid digestate for microalgae cultivation.
Liquid digestate and LF can be used more indirectly for the recovery of value-added
products by providing nutrients and other compounds (such as phytohormones) for
living organisms. This is the case for microalgae recovery (presented previously),
but also for a large range of emerging technologies. It can be highlighted (not
exhaustive): (1) Nutrient and phytohormone source for hydroponics (Antón et al.
2017; Ronga et al. 2019b; Guilayn et al. 2020a), (2) Macrophytes recovery. A full-
scale demonstration plant of duckweed recovery was recently constructed in Europe
(Pascual 2016) and (3) Liquid culture medium. For instance, liquid digestates have
been successfully used as a culture medium to produce bioplastics (Passanha et al.
2013) and ethanol (Ujor et al. 2020).
9.8.1 Composting
and solid or liquid phases (Reibel and Leclerc 2018). Due to the more favourable
thermodynamics and aerobic conditions, a higher variability of microorganisms
(i.e. bacteria, actinomycetes, fungi) can be found both effects allowing a higher
OM decomposition than AD (Guilayn et al. 2020b).
Industrial composting can be controlled by several parameters as the retention
time (from 3 weeks up to 1 or 2 months), the optimal C to N ratio between 20 and
40, the temperature, the moisture (> 60%), the bulking agent and aeration rate and
method (Epstein 2011). Composting is an exothermic process. Temperatures reach
more than 70 C within the piles inducing an authorized hygienisation process of
digestates in several countries. Temperature, exposure time and the number of
turnings are usual regulatory parameters varying from country to country. For
example, in Germany, temperatures of 55, 60 and 65 C are requested for a period
of respectively 14, 6 and 3 days (Amlinger and Blytt 2013). Largely due to the self-
heat reaction, water loss occurs during composting and decreases transportation
costs. Levasseur et al. (2017) reported a mass loss from 30 to 50%. As physical
impurities and trace metals are conserved during composting, depending on the mass
loss rate and the bulking agent used, these compounds can be either concentrated or
diluted (compared to digestate). Besides, bioaccessibility and solubility of heavy
metals can be reduced due to their strong bonds to compost organic matter (Smith
2009).
Composting is usually performed on the solid phase of digestates. Besides, some
digestates need bulking agents because of (1) the low part of biodegradable OM
allowing temperature increase, (2) the physical structure of digestates which do not
allow sufficient aeration and (3) the C to N ratio of digestates which are lower the
optimal values (Tremier et al. 2014; Zeng et al. 2014, 2015). Common bulking
agents are green waste, wood chips, sawdust and compost grinding wastes (Epstein
2011). A recent and new technique for liquid wastes is based on the spreading and
the constant turning of the liquid into a saturated support bulking material
(Chiumenti 2015; Levasseur et al. 2017).
Composting can convert ammonia to nitrates through nitrification. However,
depending on the process operation, ammonia can be largely volatilized, denitrified
(under anoxic conditions) and lost as N2 or even N2O (a strong greenhouse gas).
Zeng et al. (2012) observed 2–43% of total N loss as ammonia and up to 76%
including nitrification-denitrification. More recently, the addition of biochar has
been used as co-substrate to avoid NH3 volatilisation due to adsorption and support
for nitrifying bacteria (Wu et al. 2017). Researches indicate that industrial
composting can be optimized to produce minimal amounts of NH3 and N2O (<1%
input N) (Chiumenti 2015), but without a diffcult-to-achieve complete nitrification,
an important N2 emissions still represent a relevant loss in value.
Despite all the advantages of the composting process, composting is not neces-
sarily a low-cost technology. Moreover, the compost producer usually receives a
price near to zero or even pays an intermediary distributor. According to Dahlin et al.
(2015) and (Guilayn et al. 2020b), composts price goes from under 0 to 7 €/t, way
below other value-added products that can be recovered from digestates.
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 247
Regarding composting costs, the long retention time and the bulking agent
volume induce an installation four times or more larger than the surface needed
for AD. Furthermore, providing aeration and the subsequent air treatment for a high
surface can be extremely costly. The use of forced aeration and/or mechanical
turning is needed. Low-cost passive aeration based only on convection (heat moving
up) is usually not allowed for producing certified composts despite its attested
effectiveness. Adding to that, composting leachate treatment can be challenging
(Roy et al. 2018).
Solid digestates and solid fractions can be dried through thermal and solar drying
equipment such as those conventionally applied to wastewater sludge for several
decades. Drying equipment present a high capital investment (typically over 500 k
€), being mostly restricted to large facilities. Commonly related equipment are belt,
rotary and disk dryers. Belt driers seem more suited as rotary and disk dryers can be
damaged by digestates containing large particles. Fire risk due to self-heating and
ignition risk of explosive atmosphere must be considered during design and opera-
tion of digestate drying and storage facilities (Guilayn et al. 2020b).
From a fertilisation perspective, thermal drying induces ammoniacal nitrogen loss
through volatilisation as NH3. It can be though recovered as ammonium salt/acidic
solution through air collection and treatment (Vaneeckhaute et al. 2017). However,
in practice, many operators can be reluctant to operate air treatment equipment
(i.e. acid washing towers) in less acidic levels (e.g. pH 5–6), as any emission risk
might represent air pollution, neighbourhood complains and penalties. In such cases,
the final product can be too acidic for farmers (pH 2–3) and represent a disposal cost
superior to 40 €/m3 in Europe (SUEZ expertise). In the opposite, digestate can be
acidified prior to thermal drying to avoid nitrogen loss (Pantelopoulos et al. 2016).
By removing most of digestate moisture, thermal drying can be an effective
option for enabling large-scale facilities to export digestate to distant areas, while
maintaining product stability over time. Transportation costs can be required to be
further reduced by pelletisation, as it increases product bulk weight from 100 to over
600 kg/m3 (Dahlin et al. 2015).
experiences have already been carried out by using either the liquid or the solid
fraction of the digestate as obtained from conventionally applied separation tech-
niques (Cesaro 2021). O’Brien et al. (2019) successfully tested solid digestates
derived from dairy manure and food waste in the cultivation of the fungal species
Pleurotus ostreatus.
Waste-based composts have been extensively studied as growth media in hydro-
ponic cultivation, while the use of the digestate is more recent. The potential of the
anaerobic digestate has been also explored as growth medium in hydroponic systems
(Cesaro 2021). These can be considered as an engineered plant cultivation method,
which uses soil-less growth medium and a nutrient solution. Ronga et al. (2019b)
evaluated the cultivation of baby leaf lettuce in hydroponic systems, using both the
solid and the liquid fraction of digestate as alternative growth medium and nutrient
solution, respectively.
Cerda et al. (2019) have proposed a solid-state fermentation (aerobic) technology
based on solid food waste digestate as substrate/support medium. They have
observed interesting yields for the production of Bacillus thuringiensis (Bt), the
major biological pesticide in the organic farming market. The final product can be
used as a Bt-enriched compost-like product, or, preferentially, an extraction step
should be added to produce a further value-added extract for foliar application.
These last decades, a high interest has been brought among the scientific communi-
ties to valorise the solid fraction of digestates through thermo-chemical processes
(mainly hydrothermal, gasification and pyrolysis) (Freda et al. 2019; Miliotti et al.
2020; Cesaro 2021). Most of the studies were performed on digestate originating
from OFMSW, agricultural or WWTP sludges. Among them a specific attention has
been paid on the pyrolysis process (Torri and Fabbri 2014; Neumann et al. 2015;
Monlau et al. 2015a; Tayibi et al. 2021). Pyrolysis is defined as the thermal
decomposition of the organic matrix under non-oxidizing or very low-oxidizing
stoichiometric atmospheres, and occurs in the temperature range of 250–1200 C
(Bridgwater 2012; Biswas et al. 2017). During pyrolysis, organic matter such as
lignin, proteins, cellulose, and hemicelluloses are thermally broken down and
rearranged forming three major products (Fig. 9.3): (1) biochar (carbonaceous
solid fraction), (2) bio-oil (mainly composed of 15–30% w/w liquid of wide variety
of organic components and 85–70% of aqueous pyrolysis liquid (APL)) (Fabbri and
Torri 2016), and (3) syngas composed of non-condensable gases (i.e. CO, CO2, CH4
and H2) (Tayibi et al. 2020). Depending on the operating conditions
(i.e. temperature, heating rate, residence time) the pyrolysis process can be divided
into three main subclasses: slow, fast and flash pyrolysis (Laird et al. 2009).
Syngas can represent an interesting supplementary energy source that can be
further converted into heat or heat/electricity alone or mixed with biogas in boilers,
engines but also used for methanol production (Gollakota et al. 2016; Giuliano et al.
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 249
Fig. 9.3 The role of biochar in improving the anaerobic digestion chain
2020). Syngas has also attracted recently the attention as hydrogen source in
biological methanation processes (Schwede et al. 2017; Grimalt-Alemany et al.
2018). In parallel, bio-oil is generally separated into an organic and aqueous
phase, the first one can be used as a blend with existing transportation fossil fuels
or directly used as transportation fuels (Gollakota et al. 2016). On the contrary, the
aqueous phase can be used as new feedstock in the AD process in co-digestion with
other organic wastes (Torri and Fabbri 2014), but specific attention should be paid to
the furans and phenolic compounds present that can inhibit the microbial commu-
nities of the AD process (Barakat et al. 2012; Monlau et al. 2014).
Pyrolysis has been applied on solid digestate from OFMSW (Giuliano et al.
2020), agricultural wastes (Feng and Lin 2017; Monlau et al. 2015a; Tayibi et al.
2020), biowastes (Opatokun et al. 2017; Giwa et al. 2019) and wastewater sludges
(González-Arias et al. 2020). In most cases, slow pyrolysis was applied with
250 F. Guilayn et al.
temperatures ranging from 350 to 750 C (Neumann et al. 2015; Ghysels et al. 2020)
with most of the studies operating at 500 C (Monlau et al. 2015a; Ghysels et al.
2020; Tayibi et al. 2021). Tayibi et al. (2020) have investigated the pyrolysis
(500 C, 10 C/min, 1 h) of solid digestate treating sludges and agricultural bio-
masses and reported biochar, bio-oil and syngas yields of 37%wt, 34%wt, 29%wt
respectively. Similarly, Karaeva et al. (2021) reported biochar, bio-oil and syngas
yields of 41%wt, 31%wt, 28%wt, respectively, after pyrolysis at 550 C of solid
digestate from agricultural wastes. After pyrolysis of solid digestate, syngas energy
of 12.9 MJ/Nm3 (Tayibi et al. 2021), 9.5–15.7 MJ/Nm3 (Monlau et al. 2015a),
13.1 MJ/Nm3 (Neumann et al. 2015) were reported. The organic phase of bio-oil
generally exhibited energy yields of 23.5 MJ/kg (Tayibi et al. 2021), 28.4 MJ/kg
(Monlau et al. 2015a), 33.9 MJ/kg (Neumann et al. 2015), 24.9–32 MJ/kg (Ghysels
et al. 2020).
Biochar can be used for several applications especially for improving the stability
of composting and AD (Torri and Fabbri 2014; Fagbohungbe et al. 2017; Xiao et al.
2017), in agronomic applications alone or in combination with fertilizers like liquid
digestate (Laird et al. 2010; Semida et al. 2019; Guilayn et al. 2021; Tayibi et al.
2021), or for gas treatment like biogas purification and upgrading. A lot of synergies
have been identified in the recent years by associating AD and pyrolysis and most of
them are using biochar as a central element as shown in Fig. 9.3 (Fabbri and Torri
2016; Fagbohungbe et al. 2017; Luz et al. 2018). Up to date, even if such processes
are mature and applied at industrial scale, there is no real example that has been
tested at industrial scale for such dual symbiotic approach.
As previously mentioned, the solid digestate from the AD process especially coming
from agricultural biogas plants represents a promising perspective for the production
of biofuels such as bioethanol (Logan and Visvanathan 2019). Indeed, during the
AD process of agricultural residues such as crops and manures, part of the organic
matter is not degraded and among them a significant part of cellulose that can be
used for bioethanol production in a biorefinery concept (Ruile et al. 2015; Santi et al.
2015). Another approach for bioethanol fermentation commonly consists of two
steps: enzymatic hydrolysis and fermentation, which can be realized separately or
simultaneously. Two strategies can be generally applied: enzymatic hydrolysis
separately from fermentation, as separate hydrolysis and fermentation (SHF), or
simultaneously, as simultaneous saccharification and fermentation (SSF). In general,
the advantage of SHF is the ability to perform each step (i.e. enzymatic hydrolysis
and fermentation) under optimal temperature conditions (40–50 C for enzymatic
hydrolysis and 30 C for fermentation), whereas SSF offers the opportunities to limit
contamination and reduce the size of reactors (Cotana et al. 2015; Sukhang et al.
2020).
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 251
Recent studies have focussed on the bioethanol fermentation from solid digestate
and obtained a low ethanol yield (Monlau et al. 2015b; Sheets et al. 2015; Sambusiti
et al. 2016). Indeed, the solid digestate represents a highly recalcitrant biomass
especially due to its high lignin and crystalline cellulose content that can limit its
further biodegradation by enzymes during the saccharification process (Santi et al.
2015; Sambusiti et al. 2016). For these purposes, pretreatment technologies can be
applied such as mechanical, thermal, thermo-chemical, physical, biological or a
combination of them (Cotana et al. 2015; Sambusiti et al. 2016; Carrere et al.
2016). Until now, the pretreatments that have been reported in the literature to
enhance bioethanol fermentation from solid digestate are comprised of thermo-
acid (Teater et al. 2011; Stoumpou et al. 2020), thermo-alkaline (Yue et al. 2011;
Stoumpou et al. 2020), ozone (Wang et al. 2016) or mechanical (Sambusiti et al.
2016) treatment.
For instance, Yue et al. (2011) investigated the impact of the nature of digestates
(from CSTR and a plug flow reactor (PFR)) on bioethanol production. Ethanol
production of 105 g/kg dry digestate was noticed for the digestate from the CSTR
and 85 g/kg dry digestate for the digestate from the PFR. Similarly, Stoumpou et al.
(2020) have investigated acid and alkaline pretreatment on solid digestate (from
anaerobic digestion of wheat straw) to improve bioethanol production. Acid
pretreatment led to low bioethanol production probably due to the production of
inhibitors. In parallel, ethanol fermentation presented yields up to 65% from alkaline
pretreated digestate and all the available glucose was consumed, implying that no
inhibitory factors were present (Stoumpou et al. 2020).
From an industrial point of view, it is important that pretreatment technologies of
solid digestate take into account the ease of implementation, their cost and the
potential production of inhibitors (i.e. furans, polyphenols) (Barakat et al. 2012;
Jönsson and Martín 2016; Sukhang et al. 2020). Up to date, to our knowledge, there
is no industrial plant producing second-generation bioethanol from solid digestate.
In general, second-generation bioethanol is struggling to find industrial profitability
even if some units are already operational or under construction, most of them in
Europe and the United States. Although the first industrial-scale units appeared in the
United States (DuPont, Abengoa, Ineosbio, KiOR and Poet), some initiatives are
also existing in Europe such as the demonstrative Futurol platform (180,000 L/year)
in France and the Beta Renewables (approx. 80,000 m3/year) plant in Italy
(Crescentino).
Nutrients can be extracted from the remaining ashes after combustion of biodegrad-
able wastes. Since phosphorus and potassium are non-volatile, these macronutrients
are concentrated in the ashes (Schoumans et al. 2010). Nevertheless, these ashes can
also be rich in heavy metals, including for example cupper (Cu), cadmium (Cd) and
zinc (Zn). The available process to extract phosphorus from ashes are generally
252 F. Guilayn et al.
products derived from sewage sludge. The EPA issued a 40 CFR Rule (Part 503) that
categorizes biosolids as Class A, Class A EQ (Exceptional Quality) or Class B,
depending on the level of pathogens and the ability of the material to meet or exceed
Vector Attraction Reduction (VAR) requirements. Hereby, for Class A biosolids,
pathogens must be reduced to virtually non-detectable levels and the material must
also comply with strict standards regarding metals, odors and VAR. Class A EQ is
used to describe a biosolids product that not only meets, but exceeds, all Class A
pathogen reduction, metals and VAR requirements. Finally, Class B biosolids
contain higher levels of detectable pathogens than Class A biosolids and may require
a permit from the EPA with conditions on land application, crop harvesting and
public access. Additional state, provincial or local regulations may apply and are
currently under development for digestate and digestate-derived products.
9.10 Perspective
References
energy input and costs in cultivation using digestate and mineral fertilisation. Biomass
Bioenergy 64:199–210. https://doi.org/10.1016/j.biombioe.2014.03.061
Giuliano A, Catizzone E, Freda C, Cornacchia G (2020) Valorisation of OFMSW digestate-derived
syngas toward methanol, hydrogen, or electricity: process simulation and carbon footprint
calculation. Processes 8:526. https://doi.org/10.3390/pr8050526
Giwa AS, Xu H, Chang F, Zhang X, Ali N, Yuan J, Wang K (2019) Pyrolysis coupled anaerobic
digestion process for food waste and recalcitrant residues: fundamentals, challenges, and
considerations. Energy Sci Eng 7:2250–2264. https://doi.org/10.1002/ese3.503
Gollakota ARK, Reddy M, Subramanyam MD, Kishore N (2016) A review on the upgradation
techniques of pyrolysis oil. Renew Sustain Energy Rev 58:1543–1568. https://doi.org/10.1016/
j.rser.2015.12.180
Gong H, Yan Z, Liang KQ, Jin ZY, Wang KJ (2013) Concentrating process of liquid digestate by
disk tube-reverse osmosis system. Desalination 326:30–36. https://doi.org/10.1016/j.desal.
2013.07.010
González-Arias J, Gil MV, Fernández RÁ, Martínez EJ, Fernández C, Papaharalabos G, Gómez X
(2020) Integrating anaerobic digestion and pyrolysis for treating digestates derived from sewage
sludge and fat wastes. Environ Sci Pollut Res 27:32603–32614. https://doi.org/10.1007/s11356-
020-09461-1
Grimalt-Alemany A, Skiadas IV, Gavala HN (2018) Syngas biomethanation: state-of-the-art review
and perspectives. Biofuels Bioprod Biorefining 12:139–158. https://doi.org/10.1002/bbb.1826
Guilayn F, Jimenez J, Rouez M, Crest M, Patureau D (2017) Biogas digestate typologies (full-
paper). In: Sixteenth international waste management and landfill symposium. CISA, S.
Margherita di Pula, Cagliari, Italy. https://doi.org/10.15454/kpf7-z784
Guilayn F, Jimenez J, Martel JL, Rouez M, Crest M, Patureau D (2019a) First fertilizing-value
typology of digestates: a decision-making tool for regulation. Waste Manag 86:67–79. https://
doi.org/10.1016/j.wasman.2019.01.032
Guilayn F, Jimenez J, Rouez M, Crest M, Patureau D (2019b) Digestate mechanical separation:
efficiency profiles based on anaerobic digestion feedstock and equipment choice. Bioresour
Technol 274:180–189. https://doi.org/10.1016/j.biortech.2018.11.090
Guilayn F, Benbrahim M, Rouez M, Crest M, Patureau D, Jimenez J (2020a) Humic-like substances
extracted from different digestates: first trials of lettuce biostimulation in hydroponic culture.
Waste Manag 104:239–245. https://doi.org/10.1016/j.wasman.2020.01.025
Guilayn F, Rouez M, Crest M, Patureau D, Jimenez J (2020b) Valorisation of digestates from urban
or centralized biogas plants: a critical review. Rev Environ Sci Biotechnol 19:419–462. https://
doi.org/10.1007/s11157-020-09531-3
Guilayn F, Leurent A, Sanglier M, Rouez M, Moscoviz R, Escudie R (2021) Biochar for enhancing
anaerobic digestion of food waste. 8th International Conference on Engineering for Waste and
Biomass Valorisation, IMT-MA, France; University of Guelph, Canada, May 2021, Canada.
https://doi.org/10.15454/dcxc-7e32
Harun R, Danquah MK, Forde GM (2010) Microalgal biomass as a fermentation feedstock for
bioethanol production. J Chem Technol Biotechnol 85:199–203
Hermann L, Schaaf T (2018) Outotec (AshDec®) process for P fertilizers from sludge ash. In:
Phosphorus recovery and recycling. Springer, pp 221–233
Hidalgo D (2015) Heterotrophic microalgae cultivation to synergize anaerobic digestate treatment
with slow-release fertilizers and biostimulants production. AOP J Environ Waste Manag 1:1
Hjorth M, Nielsen AM, Nyord T, Hansen MN, Nissen P, Sommer SG (2009) Nutrient value, odour
emission and energy production of manure as influenced by anaerobic digestion and separation.
Agron Sustain Dev 29:329–338
INRA-CNRS-IRSTEA (2014) Expertise collective Mafor : Valorisation des matières fertilisantes
d’origine résiduaire sur les sols à usages agricole ou forestier
Holly et al (2017) https://doi.org/10.1016/j.agee.2017.02.007
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 257
Marcato CE, Pinelli E, Pouech P, Winterton P, Guiresse M (2008) Particle size and metal
distributions in anaerobically digested pig slurry. Bioresour Technol 99:2340–2348. https://
doi.org/10.1016/j.biortech.2007.05.013
Marcato CE, Pinelli E, Cecchi M, Winterton P, Guiresse M (2009) Bioavailability of Cu and Zn in
raw and anaerobically digested pig slurry. Ecotoxicol Environ Saf 72:1538–1544. https://doi.
org/10.1016/j.ecoenv.2008.12.010
Markou G, Wang L, Ye J, Unc A (2018) Using agro-industrial wastes for the cultivation of
microalgae and duckweeds: contamination risks and biomass safety concerns. Biotechnol Adv
36:1238–1254. https://doi.org/10.1016/j.biotechadv.2018.04.003
Markou G, Diamantis A, Arapoglou D, Mitrogiannis D, González-Fernández C, Unc A (2021)
Growing Spirulina (Arthrospira platensis) in seawater supplemented with digestate: trade-offs
between increased salinity, nutrient and light availability. Biochem Eng J 165:107815. https://
doi.org/10.1016/j.bej.2020.107815
Mazzini S, Borgonovo G, Scaglioni L, Bedussi F, D’Imporzano G, Tambone F, Adani F (2020)
Phosphorus speciation during anaerobic digestion and subsequent solid/liquid separation. Sci
Total Environ 734:139284. https://doi.org/10.1016/j.scitotenv.2020.139284
Miliotti E, Casini D, Rosi L, Lotti G, Rizzo AM, Chiaramonti D (2020) Lab-scale pyrolysis and
hydrothermal carbonisation of biomass digestate: characterisation of solid products and com-
pliance with biochar standards. Biomass Bioenergy 139:105593. https://doi.org/10.1016/j.
biombioe.2020.105593
Minasny B, Malone BP, McBratney AB, Angers DA, Arrouays D, Chambers A, Chaplot V, Chen
ZS, Cheng K, Das BS, Field DJ (2017) Soil carbon 4 per mille. Geoderma 292:59–86
Möller K (2015) Effects of anaerobic digestion on soil carbon and nitrogen turnover, N emissions,
and soil biological activity. A review. Agron Sustain Dev 35:1021–1041. https://doi.org/10.
1007/s13593-015-0284-3
Möller K, Müller T (2012) Effects of anaerobic digestion on digestate nutrient availability and crop
growth: a review. Eng Life Sci 12:242–257. https://doi.org/10.1002/elsc.201100085
Möller K, Schulz R, Müller T (2010) Substrate inputs, nutrient flows and nitrogen loss of two
centralized biogas plants in southern Germany. Nutr Cycl Agroecosyst 87:307–325. https://doi.
org/10.1007/s10705-009-9340-1
Monlau F, Sambusiti C, Barakat A, Quéméneur M, Trably E, Steyer JP, Carrère H (2014) Do
furanic and phenolic compounds of lignocellulosic and algae biomass hydrolyzate inhibit
anaerobic mixed cultures? A comprehensive review. Biotechnol Adv 32:934–951. https://doi.
org/10.1016/j.biotechadv.2014.04.007
Monlau F, Sambusiti C, Antoniou N, Barakat A, Zabaniotou A (2015a) A new concept for
enhancing energy recovery from agricultural residues by coupling anaerobic digestion and
pyrolysis process. Appl Energy 148:32–38. https://doi.org/10.1016/j.apenergy.2015.03.024
Monlau F, Sambusiti C, Ficara E, Aboulkas A, Barakat A, Carrère H (2015b) New opportunities for
agricultural digestate valorisation : current situation and perspectives. Energy Environ Sci.
https://doi.org/10.1039/C5EE01633A
Mussgnug JH, Klassen V, Schlüter A, Kruse O (2010) Microalgae as substrates for fermentative
biogas production in a combined biorefinery concept. J Biotechnol 150:51–56. https://doi.org/
10.1016/j.jbiotec.2010.07.030
Nesse AS, Sogn T, Børresen T, Foereid B (2019) Peat replacement in horticultural growth media:
the adequacy of coir, paper sludge and biogas digestate as growth medium constituents for
tomato ( Solanum lycopersicum L. ) and lettuce ( Lactuca sativa L.). Acta Agric Scand Sect B
Soil Plant Sci 69:287–294. https://doi.org/10.1080/09064710.2018.1556728
Neumann J, Binder S, Apfelbacher A, Gasson JR, García PR, Hornung A (2015) Production and
characterisation of a new quality pyrolysis oil, char and syngas from digestate – Introducing the
thermo-catalytic reforming process. J Anal Appl Pyrolysis 113:137–142. https://doi.org/10.
1016/j.jaap.2014.11.022
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 259
Nkoa R (2014) Agricultural benefits and environmental risks of soil fertilisation with anaerobic
digestates: a review. Agron Sustain Dev 34:473–492. https://doi.org/10.1007/s13593-013-
0196-z
Nyord T, Søgaard HT, Hansen MN, Jensen LS (2008) Injection methods to reduce ammonia
emission from volatile liquid fertilisers applied to growing crops. Biosyst Eng 100:235–244.
https://doi.org/10.1016/j.biosystemseng.2008.01.013
O’Brien BJ, Milligan E, Carver J, Roy ED (2019) Integrating anaerobic co-digestion of dairy
manure and food waste with cultivation of edible mushrooms for nutrient recovery. Bioresour
Technol 285:121312. https://doi.org/10.1016/j.biortech.2019.121312
Opatokun SA, Yousef LF, Strezov V (2017) Agronomic assessment of pyrolysed food waste
digestate for sandy soil management. J Environ Manag 187:24–30. https://doi.org/10.1016/j.
jenvman.2016.11.030
Pantelopoulos A, Magid J, Jensen LS (2016) Thermal drying of the solid fraction from biogas
digestate: effects of acidification, temperature and ventilation on nitrogen content. Waste Manag
48:218–226. https://doi.org/10.1016/j.wasman.2015.10.008
Panuccio MR, Papalia T, Attinà E, Giuffrè A, Muscolo A (2019) Use of digestate as an alternative
to mineral fertilizer: effects on growth and crop quality. Arch Agron Soil Sci 65:700–711.
https://doi.org/10.1080/03650340.2018.1520980
Pascual A (2016) LIFE LEMNA—duckweed technology for improving nutrient management and
resource efficiency in pig production systems. http://www.life-lemna.eu/. Accessed 21 Jan 2019
Passanha P, Esteves SR, Kedia G, Dinsdale RM, Guwy AJ (2013) Increasing
polyhydroxyalkanoate (PHA) yields from Cupriavidus necator by using filtered digestate
liquors. Bioresour Technol 147:345–352. https://doi.org/10.1016/j.biortech.2013.08.050
Reibel (2018) A Valorisation agricole des digestats: quels impacts sur les cultures, le sol et
l’environnement? Available at: https://www.geres.eu/wp-content/uploads/2019/10/ARE1805.
201.ENV_.VALOMOII.Etude_Digestats_VF.pdf
Reibel A, Leclerc B (2018) Valorisation agricole des digestats : Quels impacts sur les cultures, le sol
et l’environnement ? Revue de littérature. Méthanisation en Provence-Alpes-Côte d’Azur
Rodero Raya M, Lebrero R, Serrano león E, Lara E, Arbib Z, García-Encina PA, Muñoz R (2019)
Technology validation of photosynthetic biogas upgrading in a semi-industrial scale algal-
bacterial photobioreactor. Bioresour Technol 279. https://doi.org/10.1016/j.biortech.2019.01.
110
Ronga D, Biazzi E, Parati K, Carminati D, Carminati E, Tava A (2019a) Microalgal biostimulants
and biofertilisers in crop productions. Agronomy 9:192. https://doi.org/10.3390/
agronomy9040192
Ronga D, Setti L, Salvarani C, De Leo R, Bedin E, Pulvirenti A, Milc J, Pecchioni N, Francia E
(2019b) Effects of solid and liquid digestate for hydroponic baby leaf lettuce (Lactuca sativa L.)
cultivation. Sci Hortic 244:172–181. https://doi.org/10.1016/j.scienta.2018.09.037
Roy D, Azaïs A, Benkaraache S, Drogui P, Tyagi RD (2018) Composting leachate: characterisa-
tion, treatment, and future perspectives. Rev Environ Sci Biotechnol 17:323–349. https://doi.
org/10.1007/s11157-018-9462-5
Ruile S, Schmitz S, Mönch-Tegeder M, Oechsner H (2015) Degradation efficiency of agricultural
biogas plants—a full-scale study. Bioresour Technol 178:341–349. https://doi.org/10.1016/j.
biortech.2014.10.053
Saerens B (2017) On the benefits of phosphorus recovery from municipal wastewater. Review
paper, Aquafin NV, Belgium
Sambusiti C, Monlau F, Barakat A (2016) Bioethanol fermentation as alternative valorisation route
of agricultural digestate according to a biorefinery approach. Bioresour Technol 212:289–295.
https://doi.org/10.1016/j.biortech.2016.04.056
Santi G, Proietti S, Moscatello S, Stefanoni W, Battistelli A (2015) Anaerobic digestion of corn
silage on a commercial scale: differential utilisation of its chemical constituents and character-
isation of the solid digestate. Biomass Bioenergy 83:17–22. https://doi.org/10.1016/j.biombioe.
2015.08.018
260 F. Guilayn et al.
Schievano A, D’Imporzano G, Salati S, Adani F (2011) On-field study of anaerobic digestion full-
scale plants (Part I): an on-field methodology to determine mass, carbon and nutrients balance.
Bioresour Technol 102:7737–7744. https://doi.org/10.1016/j.biortech.2011.06.006
Schoumans OF, Rulkens WH, Oenema O, Ehlert PAI (2010) Phosphorus recovery from animal
manure. Technical opportunities and agro-economical perspectives. Alterra, Wageningen UR,
Wageningen, the Netherlands. http://content.alterra.wur.nl/Webdocs/PDFFiles/
Alterrarapporten/AlterraRapport2158.pdf
Schwede S, Bruchmann F, Thorin E, Gerber M (2017) Biological syngas methanation via
immobilized methanogenic archaea on biochar. Energy Procedia 105:823–829. https://doi.org/
10.1016/j.egypro.2017.03.396
Semida WM, Beheiry HR, Sétamou M, Simpson CR, Abd El-Mageed TA, Rady MM, Nelson SD
(2019) Biochar implications for sustainable agriculture and environment: a review. S Afr J Bot
127:333–347. https://doi.org/10.1016/j.sajb.2019.11.015
Sheets JP, Yang L, Ge X, Wang Z, Li Y (2015) Beyond land application: emerging technologies for
the treatment and reuse of anaerobically digested agricultural and food waste. Waste Manag
44:94–115. https://doi.org/10.1016/j.wasman.2015.07.037
Sialve B, Bernet N, Bernard O (2009) Anaerobic digestion of microalgae as a necessary step to
make microalgal biodiesel sustainable. Biotechnol Adv 27:409. https://doi.org/10.1016/j.
biotechadv.2009.03.001
Smith SR (2009) A critical review of the bioavailability and impacts of heavy metals in municipal
solid waste composts compared to sewage sludge. Environ Int 35:142–156. https://doi.org/10.
1016/J.ENVINT.2008.06.009
Stasinakis AS (2012) Review on the fate of emerging contaminants during sludge anaerobic
digestion. Bioresour Technol 121:432–440. https://doi.org/10.1016/j.biortech.2012.06.074
Stoumpou V, Novakovic J, Kontogianni N, Barampouti EM, Mai S, Moustakas K, Malamis D,
Loizidou M (2020) Assessing straw digestate as feedstock for bioethanol production. Renew
Energy 153:261–269. https://doi.org/10.1016/j.renene.2020.02.021
Sukhang S, Choojit S, Reungpeerakul T, Sangwichien C (2020) Bioethanol production from oil
palm empty fruit bunch with SSF and SHF processes using Kluyveromyces marxianus yeast.
Cellulose 27:301–314. https://doi.org/10.1007/s10570-019-02778-2
Takhim M, Sonveaux M, de Ruiter R (2018) The Ecophos process: highest quality market products
out of low-grade phosphate rock and sewage sludge ash. In: Phosphorus recovery and recycling.
Springer, pp 209–219
Tambone F, Genevini P, D’Imporzano G, Adani F (2009) Assessing amendment properties of
digestate by studying the organic matter composition and the degree of biological stability
during the anaerobic digestion of the organic fraction of MSW. Bioresour Technol
100:3140–3142. https://doi.org/10.1016/j.biortech.2009.02.012
Tayibi et al (2020). https://doi.org/10.1016/j.jenvman.2020.111632
Tayibi S, Monlau F, Marias F, Cazaudehore G, Fayoud NE, Oukarroum A, Zeroual Y, Barakat A
(2021) Coupling anaerobic digestion and pyrolysis processes for maximizing energy recovery
and soil preservation according to the circular economy concept. J Environ Manag 279:111632.
https://doi.org/10.1016/j.jenvman.2020.111632
Teater C, Yue Z, MacLellan J, Liu Y, Liao W (2011) Assessing solid digestate from anaerobic
digestion as feedstock for ethanol production. Bioresour Technol 102:1856–1862. https://doi.
org/10.1016/j.biortech.2010.09.099
The European Parliament and the Council of the European Union (2019) Regulation (EU) 2019/
1009 of the European Parliament and of the Council. Off J Eur Union L
Tongprawhan W, Srinuanpan S, Cheirsilp B (2014) Biocapture of CO2 from biogas by oleaginous
microalgae for improving methane content and simultaneously producing lipid. Bioresour
Technol 170:90–99. https://doi.org/10.1016/j.biortech.2014.07.094
Torri C, Fabbri D (2014) Biochar enables anaerobic digestion of aqueous phase from intermediate
pyrolysis of biomass. Bioresour Technol 172:335–341. https://doi.org/10.1016/j.biortech.2014.
09.021
9 Valorisation of Anaerobic Digestate: Towards Value-Added Products 261
Tremier A, Buffet J, Daumoin M (2014) Projet DIVA - Tâches 4.1.2 - Comportement des digestats
en compostage, 1–20
Ujor VC, Okonkwo CC, Rush BB, McCrea GE, Ezeji TC (2020) Harnessing the residual nutrients
in anaerobic digestate for ethanol fermentation and digestate remediation using Saccharomyces
cerevisiae. Fermentation 6:52. https://doi.org/10.3390/fermentation6020052
Vaneeckhaute C, Meers E, Michels E, Christiaens P, Tack FM (2012) Fate of macronutrients in
water treatment of digestate using vibrating reversed osmosis. Water Air Soil Pollut
223:1593–1603. https://doi.org/10.1007/s11270-011-0967-6
Vaneeckhaute C, Meers E, Michels E, Buysse J, Tack FM (2013) Ecological and economic benefits
of the application of bio-based mineral fertilizers in modern agriculture. Biomass Bioenergy
49:239–248. https://doi.org/10.1016/j.biombioe.2012.12.036
Vaneeckhaute C, Lebuf V, Michels E, Belia E, Vanrolleghem PA, Tack FM, Meers E (2017)
Nutrient recovery from digestate: systematic technology review and product classification.
Waste Biomass Valorisation 8:21–40. https://doi.org/10.1007/s12649-016-9642-x
Vaneeckhaute C, Belia E, Meers E, Tack FM, Vanrolleghem P (2018) Nutrient recovery from
digested waste: towards a generic roadmap for setting up an optimal treatment train. Waste
Manag 78:385–392. https://doi.org/10.1016/j.wasman.2018.05.047
Vaneeckhaute C, Darveau O, Meers E (2019) Fate of micronutrients and heavy metals in digestate
processing using vibrating reversed osmosis as resource recovery technology. Sep Purif Technol
223:81–87. https://doi.org/10.1016/j.seppur.2019.04.055
Verdi L, Kuikman PJ, Orlandini S, Mancini M, Napoli M, Dalla Marta A (2019) Does the use of
digestate to replace mineral fertilizers have less emissions of N2O and NH3? Agric For Meteorol
269–270:112–118. https://doi.org/10.1016/j.agrformet.2019.02.004
Walsh JJ, Jones DL, Edwards-Jones G, Williams AP (2012) Replacing inorganic fertilizer with
anaerobic digestate may maintain agricultural productivity at less environmental cost. J Plant
Nutr Soil Sci 175:840–845. https://doi.org/10.1002/jpln.201200214
Wang D, Xi J, Ai P, Yu L, Zhai H, Yan S, Zhang Y (2016) Enhancing ethanol production from
thermophilic and mesophilic solid digestate using ozone combined with aqueous ammonia
pretreatment. Bioresour Technol 207:52–58. https://doi.org/10.1016/j.biortech.2016.01.119
Wang Q, Prasad R, Higgins BT (2019) Aerobic bacterial pretreatment to overcome algal growth
inhibition on high-strength anaerobic digestates. Water Res 162:420–426. https://doi.org/10.
1016/j.watres.2019.07.011
Wu S, He H, Inthapanya X, Yang C, Lu L, Zeng G, Han Z (2017) Role of biochar on composting of
organic wastes and remediation of contaminated soils—a review. Environ Sci Pollut Res
24:16560–16577. https://doi.org/10.1007/s11356-017-9168-1
Xia A, Murphy JD (2016) Microalgal cultivation in treating liquid digestate from biogas systems.
Trends Biotechnol 34:264–275. https://doi.org/10.1016/j.tibtech.2015.12.010
Xiao R, Awasthi MK, Li R, Park J, Pensky SM, Wang Q, Wang JJ, Zhang Z (2017) Recent
developments in biochar utilisation as an additive in organic solid waste composting: a review.
Bioresour Technol 246:203–213. https://doi.org/10.1016/j.biortech.2017.07.090
Yue Z, Teater C, MacLellan J, Liu Y, Liao W (2011) Development of a new bioethanol feedstock—
anaerobically digested fiber from confined dairy operations using different digestion configu-
rations. Biomass Bioenergy 35:1946–1953. https://doi.org/10.1016/j.biombioe.2011.01.035
Yun Y-M, Jung K-W, Kim D-H, Oh YK, Shin HS (2012) Microalgal biomass as a feedstock for
bio-hydrogen production. Int J Hydrog Energy 37:15533–15539. https://doi.org/10.1016/j.
ijhydene.2012.02.017
Zeng Y, de Guardia A, Daumoin M, Benoist J-C (2012) Characterizing the transformation and
transfer of nitrogen during the aerobic treatment of organic wastes and digestates. Waste Manag
32:2239–2247. https://doi.org/10.1016/j.wasman.2012.07.006
Zeng Y, Guardia A, Ziebal C, De Macedo FJ, Dabert P (2014) Nitrogen dynamic and microbio-
logical evolution during aerobic treatment of digested sludge. Waste Biomass Valorisation
5:441–450. https://doi.org/10.1007/s12649-013-9275-2
262 F. Guilayn et al.
S. Wongrod (*)
Université Paris-Est, Laboratoire Géomatériaux et Environnement (EA 4508), UPEM, Marne-
la-Vallée, France
Université de Limoges, PEIRENE, Équipe Développement d0 indicateurs ou prévision de la
qualité des eaux, URA IRSTEA, Limoges, France
IHE Delft Institute for Water Education, Delft, The Netherlands
Environmental Technology Program, School of Energy, Environment and Materials, King
Mongkut’s University of Technology Thonburi, Bangkok, Thailand
Environmental and Energy Management for Community and Circular Economy (EEC&C)
Research Group, King Mongkut’s University of Technology Thonburi, Bangkok, Thailand
e-mail: [email protected]
G. Guibaud · S. Simon
Université de Limoges, PEIRENE, Équipe Développement d0 indicateurs ou prévision de la
qualité des eaux, URA IRSTEA, Limoges, France
P. N. L. Lens
IHE Delft Institute for Water Education, Delft, The Netherlands
D. Huguenot · Y. Pechaud
Université Paris-Est, Laboratoire Géomatériaux et Environnement (EA 4508), UPEM, Marne-
la-Vallée, France
E. D. van Hullebusch (*)
Université Paris-Est, Laboratoire Géomatériaux et Environnement (EA 4508), UPEM, Marne-
la-Vallée, France
IHE Delft Institute for Water Education, Delft, The Netherlands
Université de Paris, Institut de Physique Du Globe de Paris, CNRS, UMR 7154, Paris, France
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 263
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_10
264 S. Wongrod et al.
organic and inorganic pollutants in soils. This review aims to overview the produc-
tion technologies and characterization of biochars derived from sewage sludge
digestate (SSD) and the organic fraction of municipal solid waste digestate
(OFMSWD) and their applications for soil treatment. Critical discussions on the
sorption mechanism, important factors influencing contaminant retention in soils,
bioavailability of trace elements with biochar addition, and potential effects of
biochar application to soils are summarized. The interaction mechanisms involved
between the SSD and OFMSWD-derived biochars and contaminants, the main
factors influencing the biochars’ sorption efficiency and possible effects to the
environment are discussed in this review.
10.1 Introduction
The increment of food supply for human beings, rapid expansion of industrial
production and intensive agricultural activities are raising concerns. There are
many negative impacts on the environment, e.g. discharges of metal laden waste-
waters into rivers, spills of petroleum hydrocarbons to soils, transportation of
veterinary pharmaceuticals from animal manures into surface runoff and soil leach-
ate on farmland. Thus, water and soil ecosystems are prone to biomagnification and
bioaccumulation of toxic chemicals into living organisms through the food chain.
Soil contamination requires risk assessment studies and remediation to increase
environmental safety by reducing the mobility and toxicity of pollutants.
Several treatment technologies have been sought to minimize pollution (Beesley
and Marmiroli 2011), e.g. excavation, solidification and stabilization, soil washing,
phytoremediation and bioremediation (El Sawwaf and Nazir 2012; Iturbe and López
2015). However, these conventional technologies are often expensive for in situ
treatment and may lead to nutrient losses in soils (Beesley et al. 2011). Supplemen-
tation of biochar to soil is an alternative to promote environmental sustainability by
converting organic waste by-products, for instance, the solid digestate from biogas
plants, into a value-added product, i.e. biochar (Meng et al. 2013; Tan et al. 2015).
Solid digestate is an organic by-product generated by wastewater treatment
plants. Their treatment before a final disposal is required to decrease volume and
reduce unpleasant odors. Moreover, these solid digestates are considered as hazard-
ous waste due to the remaining pathogens that can possibly be transferred to the soil
(Al Seadi and Lukehurst 2012). Therefore, sludge management with environmental-
friendly and cost-effective technologies are required. Conventional technologies to
dispose sludge such as landfilling or direct use in agriculture may be restricted,
respectively, due to limited landfill site availability and possible transportation of
pollutants to the farmland (Devi and Saroha 2016). Therefore, the valorization of
digested sludge arises as an interesting approach to treat solid digestate and conse-
quently produce biochar.
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 265
Figure 10.1 shows various thermal conversion processes to produce biochar includ-
ing slow pyrolysis, fast pyrolysis, gasification, and hydrothermal carbonization
(HTC). Prior to pyrolysis or gasification, drying of the digestate to reduce the
moisture content to less than 10% is required, while no dehydration is needed for
HTC. Under slow pyrolysis, biochar is produced at low heating rates (10–30 C/min)
and long residence times (5 min–12 h) at temperatures ranging from 300 to 650 C
(Kambo and Dutta 2015). In general, higher temperatures result in lower biochar
yields due to partial degradation of lignin and cellulose (Kambo and Dutta 2015).
Table 10.1 presents the production conditions and yields of bioproducts of the
different thermal conversion technologies. The biochar produced under slow
266 S. Wongrod et al.
Fig. 10.1 Thermal conversion processes for biochar production. Adapted from Kambo and Dutta
(2015); Novotny et al. (2015); Sohi et al. (2010)
Table 10.1 Production conditions and yields of solid, liquid and gas products from thermal
conversion processes
Production conditions (operating Gas Liquid Solid
temperature, heating rate, residence (syngas) (bio-oil) (biochar)
Mode time) (%) (%) (%)
Fast pyrolysis 400–900 C, 100–1000 C/s, s–min ~13 ~75 ~12
Slow pyrolysis 250–700 C, 10–30 C/min, 15 min– ~35 ~30 ~35
2h
Hydrothermal 180–260 C, 5–10 C/min, 15 min–2 h ~5 ~25 ~70
carbonization
(HTC)
Gasification 800–950 C, 50–100 C/s, 10–20 s ~85 ~5 ~10
Adapted from Belcher and Masek (2013); Duku et al. (2011); Iwasaki et al. (2014); Kambo and
Dutta (2015); Sohi et al. (2010)
et al. 2014). High bio-oil (75%) and low biochar (10–15%) yields are often achieved
in fast pyrolysis (Table 10.1). In order to obtain homogenous biochar, pretreatment
of the feedstock, e.g. sieving to a small particle size (<2 mm) and drying (65–80 C)
to reduce the moisture to less than 10% are required (Pituello et al. 2014). Due to
lower biochar yields and higher operation costs from fast pyrolysis, the slow
pyrolysis is a more favorable option to produce biochar.
Gasification is a combination of thermal and chemical processes which is mainly
for fuel-gas production (Novotny et al. 2015). The fuel gases are mainly composed
of hydrogen (H2) and carbon monoxide (CO) that can be further used as renewable
energy sources for internal engine and power supply in industries (Novotny et al.
2015). Nevertheless, due to very high operating temperatures (800–950 C) during
gasification, the obtained biochars are relatively small, thus less desirable for biochar
production (Table 10.1). In addition, biochar produced from a hydrothermal carbon-
ization process shows low stability, i.e. high O/C ratio, whereas biochar produced
under slow pyrolysis acts as an effective sorbent due to its physicochemical
properties.
Biochar yields vary mainly depending on the operating temperature, which
strongly influences the biochar properties (Singh et al. 2010). For instance, at
increasing pyrolysis temperature (from 300 to 700 C) of SSD, a higher surface
area (from 11 to 26 m2/g) and ash content (66–87%) but lower biochar yields (from
83 to 65%) and O/C ratio (from 0.56 to 0.05; implied higher biochar stability) were
obtained (Yuan et al. 2015). In addition, the origin of feedstock also plays a role in
the biochar properties. Agrafioti et al. (2014) reported that rice husk contains lower
volatile matter (%) than sewage sludge and organic solid waste, which resulted in its
lower biochar yield.
The solid sludge by-product (solid digestate) obtained from anaerobic digestion can
be used as substrate for biochar production using pyrolysis technology (Fig. 10.2).
By coupling anaerobic digestion with the pyrolysis process, more energy recovery
and sustainability of the digestion plant were found since the bioenergy could be
used to dry the solid digestate prior to pyrolysis (Monlau et al. 2015). This chapter
focuses on two types of solid digestates: sewage sludge digestate (SSD) and organic
fraction of the municipal solid waste digestate (OFMSWD).
Sewage sludge digestate (SSD) is the organic residue remaining after the anaerobic
treatment of wastewater and waste activated sludge. The management of SSD is
currently a big concern because of the large quantities of solid digestate generated at
the wastewater treatment sites. The valorization of the solid digestate is an
268 S. Wongrod et al.
Fig. 10.2 Biochar production from organic waste digestate and its applications in soil
bioremediation
alternative to reduce the volume and eliminate organic pollutants in the digestate.
Waqas et al. (2014) reported a significant reduction of the polycyclic aromatic
hydrocarbons (PAHs) concentration after the conversion of sewage sludge into
sewage sludge-derived biochar.
Table 10.2 represents the composition of biochar derived from sewage sludge.
From Table 10.2, the SSD is rich in Cu, Cr, Pb and Zn (Pituello et al. 2014) as well as
Fe and P (Yuan et al. 2015) due to the addition of FeCl3 for phosphate precipitation
during the wastewater treatment. Moreover, the SSD biochar is rich in minerals
compared to biochars produced from biomass residues. This may result from the
diversity and complexity of sewage sludge (Zielińska et al. 2015). Sewage sludge
(59%) and its derived biochar (72%) have a high ash content, corresponding to the
dominant mineral fractions (such as SiO2, CaSO42H2O and CaCO3) contained in
the sludge (Zielińska et al. 2015). In addition, the SSD may contain undigested
organic compounds (such as carbohydrates, lipids and proteins) preliminarily from
the wastewater, inorganic elements from soils and synthetic materials from the
substrates (Devi and Saroha 2016).
Experimental P
Process for biochar conditions Yield Ash C H O N S (% or Surface area
production Tpa, RTb, HRc (%) pH (%) (%) (%) (%) (%) (%) g/kg) (m2/g) Reference
HTC 225, 16, NAd 50 NA NA 38 NA NA NA NA 6% NA Huang et al.
(2018)
Slow pyrolysis 250, 4, 3.3 69 NA NA 36 NA NA NA NA 4% NA Huang et al.
(2018)
Slow pyrolysis 250, 1, 16–19 NA 6.9 48 28 NA NA 3 0.8 17 g/kg 0.8 Pituello et al.
(2014)
Slow pyrolysis 300, NA, 10 72 5.3 52 32 2 8 3 4 NA NA Hossain et al.
(2011)
Slow pyrolysis 300, 0.5, 17 62 6.0 NA 39 4 NA 7 NA NA 4 Agrafioti et al.
(2013)
Slow pyrolysis 400, NA, 10 63 4.9 63 20 1 4 2 4 NA NA Hossain et al.
(2011)
Slow pyrolysis 450, 1, 16–19 NA 7.2 67 22 NA NA 2 0.6 21 g/kg 7 Pituello et al.
(2014)
Slow pyrolysis 450, 4, 3.3 47 NA NA 29 NA NA NA NA 7% NA Huang et al.
(2018)
Slow pyrolysis 500, 25 54 7.1 73 18 0.7 4 2 3 5% 31 Zielińska et al.
(2015)
Biochar Produced from Organic Waste Digestate and Its Potential. . .
(continued)
Table 10.2 (continued)
270
Experimental P
Process for biochar conditions Yield Ash C H O N S (% or Surface area
production Tpa, RTb, HRc (%) pH (%) (%) (%) (%) (%) (%) g/kg) (m2/g) Reference
Gasification 850, 0.57, NA 36 NA 74 21 0.5 NA 1 NA NA NA Freda et al.
(2018)
Fast pyrolysis 900, 0.033, NA 13 NA 52 33 NA 27 NA 1 3% 27 Deng et al.
(2017)
a
Tp refers to pyrolysis temperature ( C)
b
RT refers to residence time (h)
c
HR refers to heating rate ( C/min)
d
NA refers to not available
S. Wongrod et al.
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 271
significant role in the variability of the OFMSWD (Alibardi and Cossu 2015). In
contrast, Davidsson et al. (2007) showed minor effects in chemical composition of
the OFMSWD collected from different origins.
Table 10.3 shows the chemical and physical composition of biochar produced
from different OFMSWD at various production temperatures. The OFMSWD is
usually rich in Cu and Zn (Pituello et al. 2014), which further concentrate after the
anaerobic digestion of the organic waste (Tampio et al. 2016). Direct use of this
OFMSWD sludge in the field may be restricted by the presence of odors, mainly
volatile organic compounds (VOCs), and potential soil contamination by leachate
(Cesaro et al. 2019). Hence, a digestate treatment to stabilize the OFMSWD via
thermal processes (e.g. pyrolysis and gasification) (Garlapalli et al. 2016) is often
performed to eliminate the organic compounds, reduce the sludge quantities and
obtain valuable bioproducts (i.e. biochar, bio-oil and syngas).
Biochar derived from different bio-sources and pyrolyzed under several conditions
shows a wide range in chemical compositions, e.g. carbon, hydrogen, nitrogen, ash,
and metal content. The nature of the feedstock and the pyrolysis temperature are the
two most important parameters that strongly influence the chemical properties of
biochar. Low temperature biochars (300–400 C) contain several oxygen functional
groups that can interact with inorganic or polar organic chemicals through electro-
static forces and precipitation, whilst biochars produced at high temperature
(500–900 C) show high hydrophobicity which is likely to bind with organic
contaminants (Ahmad et al. 2014).
The main components of organic materials are water, hydrocarbons (HCs), and
tars (condensed polyaromatic compounds). Antal et al. (2003) demonstrated that C,
H and O contained in biochar tends to vaporize into H2O, H2, CO and CO2 with
increasing temperature. Typically, C and N are volatilized during pyrolysis, and
higher C:N ratios in biochar than in feedstocks are observed (Karaj et al. 2011;
Nartey and Zhao 2014; Yuan et al. 2015). In general, the volatilization of N and K
occurs at low temperatures, while P, S, Ca and Mg tend to mobilize at higher
temperature, and metal elements like Fe and Mn are likely to remain in the biochars
(Rodriguez 2010).
Biochar produced from SSD and OFMSWD have higher ash contents than
biochar obtained from wood and crop residues (de la Rosa et al. 2014; Pituello
et al. 2014). High ash contents in biochars can result from large mineral fractions
contained in the feedstock. de la Rosa et al. (2014) demonstrated that 69.5% of ash
was found in sewage sludge-derived biochar, whilst biochar made from wood
272
Table 10.3 Properties of biochars derived from the organic fraction of municipal solid waste digestate
Experimental P
Process for biochar conditions Yield Ash C H O N S (% or Surface area
production Tpa, RTb, HRc (%) pH (%) (%) (%) (%) (%) (%) g/kg) (m2/g) Reference
Slow pyrolysis 250, 1, 16–19 NAd 7.2 37 33 NA NA 4 1 17 g/kg 0.7 Pituello et al.
(2014)
Slow pyrolysis (mixed 300, 2, NA 92 7.8 12 40 5 53 0.3 NA 0.08 g/kg NA Rehrah et al.
material) (2015)
Slow pyrolysis 350, 1, 16–19 NA 7.6 44 34 NA NA 4 0.9 20 g/kg 5 Pituello et al.
(2014)
Slow pyrolysis 400, 0.33, NA NA 8.0 6 48 12 31 1 0.1 NA 20 Jin et al. (2014)
Slow pyrolysis 450, 1, 16–19 NA 7.4 55 29 NA NA 3 0.8 24 g/kg 27 Pituello et al.
(2014)
Slow pyrolysis 500, 0.33, NA NA 8.5 9 59 9 20 1 0 NA 29 Jin et al. (2014)
Slow pyrolysis 550, 1, 16–19 NA 7.1 61 26 NA NA 2 0.6 26 g/kg 77 Pituello et al.
(2014)
Slow pyrolysis 600, 0.33, NA NA 9.0 6 70 8 13 1 0.1 NA 29 Jin et al. (2014)
a
Tp refers to pyrolysis temperature ( C)
b
RT refers to residence time (h)
c
HR refers to heating rate ( C/min)
d
NA refers to not available
S. Wongrod et al.
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 273
contained an approximately 6 times (10.6%) lower ash content. The main composi-
tion of ash is alkaline salts of Na, K, Ca and Mg (de la Rosa et al. 2014). Many
studies showed that the percentages of ash in biochars increased along with higher
temperature, this may result from the higher organic matter volatilization with
increasing temperatures (Cantrell et al. 2012; Tsai et al. 2012; Zielińska et al.
2015). Additionally, the presence of Ca, Fe and Al oxides on the biochar solid
matrix can be attributed to the precipitation of target pollutants on the biochar
surface. The studies of Agrafioti et al. (2014) showed that biochars produced from
sewage sludge and the organic fraction of municipal solid waste contained much
higher mineral fractions (i.e. CaO, Fe2O3 and Al2O3) than biochar produced from
rice husk (<1.3% w/w). CaO was found in both SS (17% w/w) and OFMSW (50%
w/w) biochars, whereas Fe2O3 and Al2O3 were mainly found in the SS biochar (13%
and 5% w/w, respectively), but were negligible in the OFMSW biochar (< 1.5 % w/
w) (Agrafioti et al. 2014).
The physical characteristics of biochar, such as BET (Brunauer, Emmett and
Teller) specific surface area (SBET) and the crystallographic structure, also play a role
in adsorption of contaminants as well as retention of water and nutrients. Rodriguez
(2010) stated that biochars have a relatively large SBET and high porosity, which
vary depending on the type of organic material and the production conditions.
Typically, micropores enhance the sorption capacity of contaminants while
mesopores are mostly related to adsorption between solid and liquid phases, whereas
macropores are significant for the bulk soil structure and aeration in soils. However,
a higher SBET may not always corresponds to a better sorption for metal elements.
Agrafioti et al. (2014) showed that biochar produced from organic digestates have a
higher sorption efficiency for As(V) and Cr(III) than rice husk-derived biochar even
they have a much lower SBET (10–30 times) than the latter. This suggests that the
sorption mechanism of trace elements onto these biochars is predominated by
electrostatic interactions (i.e. anionic and cationic trace elements attraction) rather
than physical adsorption. Nevertheless, a high SBET and pore volume can enhance a
faster mass transfer between contaminants and biochar pores (Agrafioti et al. 2014).
Biochars are recently applied as media for retention of organic and inorganic
contaminants in soils. However, a low sorption capacity has been found in certain
conventional biochars and thus a modification post-treatment to improve the sorp-
tion efficiency for target pollutants becomes necessary. Several pretreatment
methods aiming to improve the surface charges of biochar can be applied, such as
treatments with acidic or basic chemicals, steam activation, magnetization, and
surfactant modifications (Rajapaksha et al. 2016). The selection of the treatment
method for biochar depends on the type of contaminants such as hydrocarbons
compounds, metal(loid) elements, anionic and cationic ions and the hydrophobicity
and polarity of the biochar (Rajapaksha et al. 2016).
274 S. Wongrod et al.
Activation of biochar with steam was found to have higher nutrient retention
times available for plants than the raw biochar and activated biochar showed almost
doubled positive effects on the soil than non-activated biochar (Borchard et al.
2012). Regarding the difficulty in removing powdered biochars from aqueous
solution after the treatment, magnetic modification of biochar is an interesting
technique. Devi and Saroha (2014) found that zero valent iron (ZVI) impregnated
biochar composites, synthesized by FeSO47H2O as a chemical reagent and NaBH4
as a reducing agent, provided simultaneous adsorption and dechlorination of
pentachlorophenol.
Chemical modifications of biochars can be carried out by treatment with acidic
chemicals such as phosphoric acid (H3PO4), sulfuric acid (H2SO4) and nitric acid
(HNO3), an oxidizing agent such alkaline chemicals such as sodium hydroxide
(NaOH) and potassium hydroxide (KOH) (Rajapaksha et al. 2016). H3PO4 is one
of the common reagents used to modify chemical structures of adsorbents (Kang
et al. 2011). Treating with H3PO4 helps to break down lignocellulose, aliphatic and
aromatic biomass and further forming phosphate and polyphosphate to avoid con-
traction during the development of pores (Yang et al. 2011). H2O2, as an oxidant, has
been used to modify biochar and induce more oxygen-containing functional groups,
particularly carboxylic groups, on its surface (Gao et al. 2012). In addition, KOH
was found to induce higher SBET and more available functional groups on the
biochar surface (Jin et al. 2014). Moreover, biochar treated with urea had more
available N functional groups and surface basicity (Stavropoulos et al. 2008).
Regarding the improved properties of biochars after modification, the sorption
capacity of the modified biochars towards both inorganic and organic pollutants
have been investigated. The Pb sorption capacity was enhanced from 0.88 mg/g (raw
biochar) to 22.82 mg/g on H2O2-modified biochar, in accordance with more carbox-
ylic groups available on the biochar surface (Gao et al. 2012). Modification of
municipal solid waste-derived biochar with KOH was found to enhance the sorption
efficiency of As(V) 1.3 times, in agreement with its higher SBET and more available
surface functional groups (Jin et al. 2014). Rajapaksha et al. (2014) stated that
removal of organic pollutants could be improved by the π-π electron donor interac-
tions of biochar with the aromatic rings of organic pollutants. Biochar treated with
urea was also found to have higher uptake rates of phenol due to better π-π dispersion
interactions (Stavropoulos et al. 2008).
10.4 Applications
Figure 10.3 shows the main mechanisms of sorption for organic and inorganic
pollutants onto the biochar surface. The main mechanisms between biochar and
organic pollutants are electrostatic forces, active binding of oxygen functional
groups on biochar pores with polar organic compounds and bonding of non-polar
organic compounds on hydrophobic sites of biochar (Fig. 10.3). In general, organic
waste-derived biochar produced at high pyrolysis temperatures possesses a higher
sorption ability for hydrocarbon compounds due to their large SBET, high porosity
and high hydrophobicity on the biochar surface.
Sorption mechanisms between biochar and metal elements include ion exchange
between these elements and Na, Ca or K present on the biochar surface, electrostatic
attractions with the available surface functional groups (e.g. carboxyl and hydroxyl)
276 S. Wongrod et al.
Fig. 10.3 Interactions of biochar with positively and negatively charged pollutants
and co-precipitation of metals with oxide groups of biochar and humic substances
(Ahmad et al. 2014; Inyang et al. 2016; Zhang et al. 2013).
The adsorption and partitioning of organic pollutants in soils are also affected by
the clay minerals and the organic carbon content present in the native soil (Ahmad
et al. 2014; Zhang et al. 2015). Zielińska and Oleszczuk (2016) found that the
presence of dissolved organic carbon (DOC) and clay minerals in the soil play
also a role in reducing the sorption ability of phenanthrene and pyrene by sewage
sludge-derived biochar. Zhang et al. (2015) investigated that with the same biochar
content, a low organic carbon content in the soil exhibited a 10 times higher
adsorption coefficient (Kf) than soil with a high organic carbon content, and
98 times higher compared to the control soil. This shows that biochar enhances
more the sorption ability of soils with a low organic carbon content.
Table 10.4 shows the sorption characteristics of organic waste digestate-derived
biochar for organic and inorganic contaminants. With different composition in the
feedstock, biochar can behave differently in the sorption interaction (Table 10.4).
The mobility of Cu is significantly affected by the organic carbon content in biochar
(Beesley et al. 2010). Yang et al. (2016) studied the effect of coexisting Al on P
sorption in acidic polluted soils upon the addition of biochar made from sewage
sludge. Al was shown to reduce the on the biochar via pH buffering against biochar
alkalinity (Yang et al. 2016). They also reported the enhancement of Pb-K
co-precipitation at lower pH and coexisting Al, whilst the co-precipitation of Pb-P
on sewage sludge biochar was the dominant Pb sorption mechanism at higher pH
(Yang et al. 2016).
Table 10.4 Sorption characteristics for organic and inorganic contaminants on biochar produced from organic waste digestate
Isotherm/
CRa, kinetic Adsorption
Biochar feedstock Pollutant pH contact time models capacity Main mechanisms Reference
Municipal solid As(V) 6.0 5–400 mg/ Langmuir/ 24.49 mg/g Surface complexation with O-contained Jin et al.
waste digestate L, 30 h PSOb functional groups (e.g. carboxylic and phe- (2014)
(500 C) nolic); physical sorption via surface area and
Municipal solid As(V) 6.0 5–400 mg/ Langmuir/ 30.98 mg/ pore volume of biochar
waste digestate + L, 30 h PSO g
KOH
(500 C)
Municipal organic As(V) 6.0–7.7 90 μg/L, Freundlich/ 3.54 μg/g Adsorption of As(V) with mineral fractions Agrafioti
waste digestate 72 h PSO of biochar (i.e. CaO, Fe2O3 and Al2O3); et al.
(300 C) Cr(III) 7.0–9.5 170 μg/L, Freundlich/ 42.37 μg/g electrostatic interactions between negative (2014)
72 h PSO surface charge of biochar and Cr(III) cations
Sewage sludge As(V) 6.0–7.7 90 μg/L, Freundlich/ 4.25 μg/g
(300 C) 72 h PSO
Cr(III) 7.0–9.5 170 μg/L, Freundlich/ 30.12 μg/g
72 h PSO
Waste-art-paper Pb 5.0 20–400 mg/ Langmuir/ 51.2 mg/g Precipitation of Pb with carbonate to form Xu et al.
digestate (300 C) L, 24 h PSO PbCO3 and Pb2(OH)2CO3 on biochar (2017)
surfaces
Municipal sewage Methylene blue 2.0–11.0 50–150 mg/ Langmuir/ 12.58 mg/g Endothermic adsorption; electrostatic inter- Fan et al.
sludge + tea waste L, 24 h PSO action; ion exchange; surface complexation (2016)
(300 C)
Paper mill sludge Tricaine 7.0–7.5 10–500 mg/ Freundlich/ 87.4 mg/g The p–p interactions between aromatic Ferreira
(800 C) methanesulfonate L, 500 min PSO structure of tetracycline and graphite-like et al.
sheets of biochar (2017)
a
CR refers to concentration range
b
PSO refers to pseudo-second-order
S. Wongrod et al.
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 279
In recent studies, biochar produced from digested organic sludge has been recog-
nized as a potential tool for the sorption of organic pollutants in soils (Khan et al.
2013; Waqas et al. 2014; Zielińska and Oleszczuk 2015). The presence of micro-
pores and mesopores on the biochar surface as well as aromatic structures and
functional groups help to retain toxic contaminants like PAHs, atrazine and
sulfamethazine in soils (Cao et al. 2011; Rajapaksha et al. 2014; Vithanage et al.
2014). Soil amended with biochar reduced the dissolved concentrations of phenan-
threne and pyrene compared to soils without biochar addition (Zielińska and
Oleszczuk 2016). Addition of 5% w/w sewage sludge-derived biochar to soil
enhanced the sorption capacity of phenanthrene from 8.3 to 20.3% (Zielińska and
280 S. Wongrod et al.
Oleszczuk 2016). Also, the sorption of pyrene in soil amended with sewage sludge-
derived biochar was increased from 14.5 to 31.7% with respect to control soil
(Zielińska and Oleszczuk 2016). Mukherjee et al. (2016) investigated the potential
adsorption-desorption of pesticides (i.e. bentazone, boscalid and pyrimethanil) using
digestates from a biogas plant (i.e. maize silage, manures and urine) in a loamy sand
soil. The partition coefficient (Koc) values of the pesticide boscalid were approxi-
mately 5 times higher in soil mixed with 30% digestate and 5% biochar compared to
the control soil.
Application of sewage sludge and its derived biochar to soils significantly
decreased the available PAHs concentration (Waqas et al. 2014). Also, supplemen-
tation of sludge and sludge-based biochar decreased the bioaccumulation of PAHs
into Cucumis sativa L. by 44–57% with biochar amendment and 20–36% with
sewage sludge amendment (Waqas et al. 2014). These results show that biochar
promotes a higher efficiency in the immobilization of PAHs than the raw sewage
sludge digestate during soil remediation. However, a higher uptake rate of
terbuthylazine in biochar amended in low organic carbon soil was found compared
to soil with a high organic carbon content. This suggests that the presence of
dissolved organic carbon in soil may block the biochar pores, resulting in lower
accessibility of pesticides to biochar (Ahmad et al. 2014).
Table 10.5 Application of organic waste digestate-derived biochar for organic and inorganic
contaminant remediation
Pyrolysis
Inorganic Biochar temperature
contaminant feedstock ( C) Effect Reference
As, Cd, Cr, Co, SSD 500–550 Immobilization of As, Cr, Co, Ni Khan et al.
Cu, Ni, Pb and and Pb (rise in pH) but mobiliza- (2013)
Zn tion of Cu, Zn, and Cd in soils
Cu, Cd, Ni, Zn SSD 500 Biochar increased soil respiration Méndez
and Pb compared to control soil but lower et al.
rate than soil with SSD. Lower (2012)
leaching of Cu, Ni and Zn in soil-
biochar amendment than in soil
amended with SSD
Cu, Pb, Zn, Cd SSD 450 Biochar enhanced soil fertility and Liu et al.
and Cr crop productions. Reduction of (2014)
metals availability for plant which
decrease the toxicity in contami-
nated soil
As, Cd, Cu, Pb SSD 550 Soil + SSD biochar: significantly Waqas
and Zn decreases of As, Cu and Pb con- et al.
centrations (2014)
Soil + SSD amendment: increases
in metal concentrations
Phenanthrene SSD 500 and 700 Application of 5% w/w biochar Zielińska
(PHE) and enhanced 2.4- and 2.2-times and
pyrene (PYR) higher sorption capacity for PHE Oleszczuk
and PYR, respectively in soils. (2016)
The sorption increased with
biochar produced at higher tem-
perature. DOC and clay minerals
present in the soils reduced the
efficiency of PHE and PYR sorp-
tion by biochar
Polycyclic aro- SSD 550 PAH concentrations in biochar Waqas
matic hydrocar- significantly decreased after et al.
bons (PAHs) pyrolysis. Biochar addition to soil (2014)
reduced availability and
bioaccumulation of PAHs into
Cucumis sativa L.
Halogenated OFMSWD 400 The sorption efficiency of biochar Oh and
phenols effectively increased at pH 4 or Seo (2016)
7. Surface area and carbon content
of biochar and hydrophobicity of
halogenated phenols mainly affect
the phenol sorption
Pharmaceuticals OFMSWD 400 pH significantly affected the Oh and
(i.e. ibuprofen) sorption capacities of biochar for Seo (2016)
ibuprofen due to the point of zero
charge (PZC) of sorbents and
deprotonation of the
pharmaceutical
282 S. Wongrod et al.
The mobility and fate of metals in contaminated soil amended with biochar have
been widely studied (Beesley et al. 2014; Houben et al. 2013; Kargar et al. 2015;
Nartey and Zhao 2014). The effects on metal bioavailability in soils vary depending
on the nature of the biochar feedstock and type of metal (Zhang et al. 2013).
Addition of organic wastes as soil amendments to contaminated soils can increase
the bioavailability of metal(loid)s (Nartey and Zhao 2014). Tampio et al. (2016)
reported that plants had a higher metal contents and lower N bioavailability upon
amendment of a soil with a mixture of vegetable waste digestate and activated-
sludge. However, addition of biogas digestate from food waste and the organic
fraction of municipal solid waste to the soil showed positive outcomes as they
provided more bioavailable nutrients in the soil (Tampio et al. 2016).
Biochar produced from municipal wastewater treatment sludge contains mineral
oxides and a carbonized organic fraction which serves as a good sorbent for metals
(Zhang et al. 2016). The precipitation of metal (Pb and Zn) phosphate and more
available sorption sites were observed after aging of biochar for 30 days, ensuring
durability of metal immobilization in sludge-derived biochar during soil amendment
(Zhang et al. 2016). Campos and De la Rosa (2020) reported that Cu and Pb were
effectively immobilized in soils amended with biochar, which affirmed the promis-
ing potential of biochar to recover land polluted with heavy metals. It is important to
understand that one type of biochar cannot be representatively used as sorbent to
remediate all types of metals. Hence, the use of solid digestate-derived biochar in
soils should be verified for each particular biochar type to ensure the environmental
safety for field application.
lower O:C ratio and biochar with an O:C ratio lower than 0.2 shows a half-life of
more than a thousand year, whilst a reduction of the biochar half-life to less than a
hundred year has been observed with an O:C ratio higher than 0.6 (Kuppusamy et al.
2016). Also, the pyrolysis temperature and feedstock source are two main parame-
ters affecting the O and C content of a biochar (Spokas 2010). The O:C ratio is likely
to provide a strong indicator of biochar stability in soil rather than solely the
operation conditions or feedstock type (Spokas 2010). In addition, the biochar
persistence in soil can be predicted from the fraction of aromatic rings and
non-aromatic compounds present in the biochar (Singh et al. 2012). Hernandez-
Soriano et al. (2016) observed that addition of biochar to loamy soil slowed down the
soil respiration rate compared to soils amended with raw organic residues during
237 days of incubation. These results showed that the carbon metabolism in soil
amended with biochar enhanced the carbon stability in the soil (Hernandez-Soriano
et al. 2016). After 4 years of soil-biochar incubation in a field experiment, a
significant increase of the C:N ratio of the microbial biomass was found in
biochar-amended soil (Zhang et al. 2014). Additionally, they suggested that biochar
can reduce the variability of the environmental conditions for microorganisms and
thus decrease the temporary fluctuations in C and N dynamics.
Apart from the biochar applications as soil amendments, the use of biochar as
active sorbent to retain contaminants in soils in a long term has also been reported.
Liu et al. (2019) found that after 240 days of sorption equilibration of lincomycin
(an antibiotic) by biochar, about 75% of sorbed lincomycin on the biochars could not
be extracted anymore with acetonitrile/methanol extractant. This suggests that
biochar has a great potential to retain antibiotics in a long term. Moreover, Stefaniuk
et al. (2017) found a significant decrease in the PAHs concentrations after a long-
term incubation of sewage sludge and/or biochar in the soil (1.5 years), which were
reduced by 19% upon the addition of sewage sludge and by 35% with the addition
(5.0%) of sewage sludge and biochar. He et al. (2019) found that a soil incubation
with sewage sludge derived biochar can immobilize Pb and Cu in a long term
(2 years).
pore water (Beesley et al. 2011). Biochar added in agricultural soils can provide
positive impacts on plant growth such as maize (Agegnehu et al. 2016; Gwenzi et al.
2016) due to the liming effect and a longer nutrient retention time in the soils (Sohi
et al. 2010). Conversely, adverse effects on plant growth were found by van Zwieten
et al. (2010) since biochar-soil amendment induced reduction of biomass production
by wheat and radish.
Earthworms serve as ecosystem engineers as they play important roles in soils,
i.e. they accelerate organic matter degradation, increase nutrient availability to plants
and nutrient recycling in ecological systems (Beesley et al. 2011). Therefore,
earthworms are considered as vital soil animals for which a reduction of contami-
nants becomes significant to avoid an over-uptake (Sizmur et al. 2015). From
Table 10.6, addition of sewage sludge-derived biochar in a mixture of sewage sludge
and wood chips integrated with vermicomposting showed 11- and 5-times higher
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 285
Soil quality deterioration is a worldwide problem and has been particularly observed
in agricultural areas. Addition of organic amendments to soils may be beneficial to
restore the soil quality. Addition of labile organic materials to soil can provide
bioavailable nutrients for crops and soil microbiota like microbes and fungi. How-
ever, a low nitrogen content in the organic substrates can limit the plant growth (Sohi
et al. 2010). Supplementation of alkaline biochar to acidic soils can enhance crop
yields through increasing soil pH and reducing acidity constraints which inhibit the
plant productivity (Ahmad et al. 2014). The efficiency of biochar amendment
depends on the biochar buffering capacity that varies with the nature of feedstock
and the pyrolytic temperature (Ahmad et al. 2014; Tables 10.2 and 10.3).
The use of biochar, compost and their combination in a tropical agricultural soil
has been studied: significant increases in total N, available P, nitrate nitrogen and
ammonium nitrogen in biochar-compost-soil systems compared to the control were
found (Agegnehu et al. 2016). Besides, biochar mixed with compost increased crop
biomass by 10–29% and helped to reduce N2O emissions compared to other
treatments (Agegnehu et al. 2016). Gwenzi et al. (2016) reported that with the
application of sewage sludge mixed with its derived biochar (15 or 7.5 t/ha of
each organic amendment) on tropical clay soil showed a 47% enhancement of
maize growth and was comparable to inorganic fertilizer (49%). Additionally,
adding biochar and sewage sludge in the tropical clay soil reduced Pb, Cu and Zn
uptake by the maize plants by 22% with respect to the presence of sludge alone
during the treatment (Gwenzi et al. 2016).
Supplementation of 12% biochar significantly reduced total nitrogen losses and
GHGs emissions, i.e. CH4, N2O and NH3, respectively by 92.85, 95.14 and 58.03%
during sewage sludge composting (Awasthi et al. 2016). However, low dosage
(2–6%) of biochar added to compost released more CH4 and N2O than the control
treatment (Awasthi et al. 2016). In addition, the humic (42%) and fulvic (28%) acid
concentrations increased during biochar-compost treatment with 12% biochar addi-
tion (Awasthi et al. 2016). van Zwieten et al. (2010) applied paper mill waste sludge-
derived biochar in two different agricultural soils, i.e. a ferrosol (high agricultural
potential) and calcarosol (low-moderate agricultural potential) in a greenhouse
study. The results showed high N uptake rates by wheat crops in biochar-ferrosol
amended soil and 250% increases in biomass production of soybean and radish
286 S. Wongrod et al.
compared to the control (van Zwieten et al. 2010). However, the calcarosol showed
varied results, i.e. increases in soybean biomass but decreases of wheat and radish
biomass (van Zwieten et al. 2010).
Biochar application to soil can alter microbial activities since the degradation of soil
organic carbon provides energy for microbes, with addition of hardly degradable
carbon (i.e. biochar) to the soil. Thus, possible adverse effects on the microbial
community in soil can occur, generally depending on the types of organic residues as
feedstock, biochar dosage and time interval of its application (Sohi et al. 2010).
Weyers and Spokas (2011) reported a reduction of earthworm biomass when organic
sludge-derived biochars were added to soils.
It is also important to consider priming effects when biochar is applied to soil.
The priming effect is a phenomenon occurring when higher rates of soil organic
carbon decomposition take place corresponding to more organic materials
supplemented to the soil. This enhances the activity of the microbial population
due to the release of energy during the decomposition of organic substrates, which
accelerates the mineralization of soil organic carbon (Fontaine et al. 2003). Positive
priming effects were observed by faster decomposition rates of soil carbon when
amended with biochar, resulting in microbial activity enhancement and more hydro-
lysis reactions, which effectively increase with higher soil pH (Ahmad et al. 2014;
Kuzyakov et al. 2009). In contrast, a negative priming effect can occur when biochar
adsorbs dissolved organic carbon in soils, which lowers the decomposition rates and
subsequently reduces the microbial population.
Short-term experiments (less than a few years) confirm the potential use of organic
sludge biochars as soil amendments and sorbing materials in the environment.
However, based on pot experiments in small scale, more studies on field sites are
necessary to confirm if similar outcomes are achieved. Also, long-term studies of
sludge biochar application in soils are lacking and yet to be further developed. The
studies on the longevity of microbial activities in soil-biochar systems and the
durability effect of organic sludge biochar in the long-term should be further
achieved as well as the regeneration of aged biochar.
Acknowledgements This chapter emanated from the PhD thesis entitled “Biochars from solid
digestates as sorbing materials for metal(loid)s removal from water” by Suchanya Wongrod. The
authors are thankful to the financial support by the Marie Sklodowska-Curie European Joint
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 287
References
Agegnehu G, Bass AM, Nelson PN, Bird MI (2016) Benefits of biochar, compost and biochar-
compost for soil quality, maize yield and greenhouse gas emissions in a tropical agricultural soil.
Sci Total Environ 543:295–306
Agrafioti E, Bouras G, Kalderis D, Diamadopoulos E (2013) Biochar production by sewage sludge
pyrolysis. J Anal Appl Pyrolysis 101:72–78
Agrafioti E, Kalderis D, Diamadopoulos E (2014) Arsenic and chromium removal from water using
biochars derived from rice husk, organic solid wastes and sewage sludge. J Environ Manag 133:
309–314
Ahmad M, Rajapaksha AU, Lim JE, Zhang M, Bolan N, Mohan D, Vithanage M, Lee SS, Ok YS
(2014) Biochar as a sorbent for contaminant management in soil and water: a review.
Chemosphere 99:19–23
Alibardi L, Cossu R (2015) Composition variability of the organic fraction of municipal solid waste
and effects on hydrogen and methane production potentials. Waste Manag 36:147–155
Al Seadi T, Lukehurst CT (2012) Quality management of digestate from biogas plants used as
fertiliser. IEA Bioenergy
Antal MJ, Mochidzuki K, Paredes LS (2003) Flash carbonization of biomass. Ind Eng Chem Res
42:3690–3699
Arazo RO, Genuino DAD, de Luna MDG, Capareda SC (2017) Bio-oil production from dry sewage
sludge by fast pyrolysis in an electrically-heated fluidized bed reactor. Sustain Environ Res 27:
7–14
Awasthi MK, Wang M, Chen H, Wang Q, Zhao J, Ren X, Li D, Awasthi SK, Shen F, Li R, Zhang Z
(2016) Heterogeneity of biochar amendment to improve the carbon and nitrogen sequestration
through reduce the greenhouse gases emissions during sewage sludge composting. Bioresour
Technol 224:428–438
Beesley L, Marmiroli M (2011) The immobilisation and retention of soluble arsenic, cadmium and
zinc by biochar. Environ Pollut 159:474–480
Beesley L, Moreno-Jiménez E, Gomez-Eyles JL, Moreno-Jimenez E (2010) Effects of biochar and
greenwaste compost amendments on mobility, bioavailability and toxicity of inorganic and
organic contaminants in a multi-element polluted soil. Environ Pollut 158:2282–2287
Beesley L, Moreno-Jiménez E, Gomez-Eyles JL, Harris E, Robinson B, Sizmur T (2011) A review
of biochars’ potential role in the remediation, revegetation and restoration of contaminated soils.
Environ Pollut 159:3269–3282
Beesley L, Inneh OS, Norton GJ, Moreno-Jimenez E, Pardo T, Clemente R, Dawson JJC (2014)
Assessing the influence of compost and biochar amendments on the mobility and toxicity of
metals and arsenic in a naturally contaminated mine soil. Environ Pollut 186:195–202
Belcher CM, Masek O (2013) 16. Biochar and carbon sequestration. In: Belcher CM (ed) Fire
phenomena and the earth system: an interdisciplinary guide to fire science. Wiley, Oxford
Borchard N, Wolf A, Laabs V, Aeckersberg R, Scherer HW, Moeller A, Amelung W (2012)
Physical activation of biochar and its meaning for soil fertility and nutrient leaching—a
greenhouse experiment. Soil Use Manag 28:177–184
Campos P, De la Rosa JM (2020) Assessing the effects of biochar on the immobilization of trace
elements and plant development in a naturally contaminated soil. Sustain 12:1–19
Cantrell KB, Hunt PG, Uchimiya M, Novak JM, Ro KS (2012) Impact of pyrolysis temperature and
manure source on physicochemical characteristics of biochar. Bioresour Technol 107:419–428
288 S. Wongrod et al.
Cao X, Ma L, Liang Y, Gao B, Harris W (2011) Simultaneous immobilization of lead and atrazine
in contaminated soils using dairy-manure biochar. Environ Sci Technol 45:4884–4889
Cesaro A, Conte A, Belgiorno V, Siciliano A, Guida M (2019) The evolution of compost stability
and maturity during the full-scale treatment of the organic fraction of municipal solid waste. J
Environ Manag 232:264–270
Chen B, Zhou D (2008) Transitional adsorption and partition of nonpolar and polar aromatic
contaminants by biochars of pine needles with different pyrolytic temperatures. Environ Sci
Technol 42:5137–5143
Chun Y, Sheng G, Chiou CT, Xing B (2004) Compositions and sorptive properties of crop residue-
derived chars. Environ Sci Technol 38:4649–4655
Davidsson Å, Gruvberger C, Christensen TH, Hansen TL, Jansen JLC (2007) Methane yield in
source-sorted organic fraction of municipal solid waste. Waste Manag 27:406–414
de la Rosa JM, Paneque M, Miller AZ, Knicker H (2014) Relating physical and chemical properties
of four different biochars and their application rate to biomass production of Lolium perenne on
a Calcic Cambisol during a pot experiment of 79 days. Sci Total Environ 499:175–184
Deng S, Tan H, Wang X, Yang F, Cao R, Wang Z, Ruan R (2017) Investigation on the fast
co-pyrolysis of sewage sludge with biomass and the combustion reactivity of residual char.
Bioresour Technol 239:302–310
Devi P, Saroha AK (2014) Synthesis of the magnetic biochar composites for use as an adsorbent for
the removal of pentachlorophenol from the effluent. Bioresour Technol 169:525–531
Devi P, Saroha AK (2016) Utilization of sludge based adsorbents for the removal of various
pollutants: a review. Sci Total Environ 578:16–33
Ding W, Dong X, Ime IM, Gao B, Ma LQ (2014) Pyrolytic temperatures impact lead sorption
mechanisms by bagasse biochars. Chemosphere 105:68–74
Duku MH, Gu S, Hagan EB (2011) Biochar production potential in Ghana—A review. Renew
Sustain Energy Rev 15:3539–3551
El Sawwaf M, Nazir AK (2012) The effect of deep excavation-induced lateral soil movements on
the behavior of strip footing supported on reinforced sand. J Adv Res 3:337–344
Fan S, Tang J, Wang Y, Li H, Zhang H, Tang J, Wang Z, Li X (2016) Biochar prepared from
co-pyrolysis of municipal sewage sludge and tea waste for the adsorption of methylene blue
from aqueous solutions: kinetics, isotherm, thermodynamic and mechanism. J Mol Liq 220:
432–441
Ferreira CIA, Calisto V, Otero M, Nadais H, Esteves VI (2017) Removal of tricaine
methanesulfonate from aquaculture wastewater by adsorption onto pyrolysed paper mill sludge.
Chemosphere 168:139–146
Florence TM, Morrison GM, Stauber JL (1992) Determination of trace element speciation and the
role of speciation in aquatic toxicity. Sci Total Environ 125:1–13
Fontaine S, Mariotti A, Abbadie L (2003) The priming effect of organic matter: a question of
microbial competition? Soil Biol Biochem 35:837–843
Freda C, Cornacchia G, Romanelli A, Valerio V, Grieco M (2018) Sewage sludge gasification in a
bench scale rotary kiln. Fuel 212:88–94
Gadepalle VP, Ouki SK, Van Herjwijen R, Hutchins T (2007) Immobilization of heavy metals in
soil using natural and waste materials for vegetation establishment on contaminated sites. Soil
Sediment Contam 16:233–251
Garlapalli RK, Wirth B, Reza MT (2016) Pyrolysis of hydrochar from digestate: effect of hydro-
thermal carbonization and pyrolysis temperatures on pyrochar formation. Bioresour Technol
220:168–174
Gwenzi W, Muzava M, Mapanda F, Tauro TP (2016) Comparative short-term effects of sewage
sludge and its biochar on soil properties, maize growth and uptake of nutrients on a tropical clay
soil in Zimbabwe. J Integr Agric 15:1395–1406
He E, Yang Y, Xu Z, Qiu H, Yang F, Peijnenburg WJGM, Zhang W, Qiu R, Wang S (2019) Two
years of aging influences the distribution and lability of metal(loid)s in a contaminated soil
amended with different biochars. Sci Total Environ 673:245–253
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 289
Hernandez-Soriano MC, Kerré B, Kopittke PM, Horemans B, Smolders E (2016) Biochar affects
carbon composition and stability in soil: a combined spectroscopy-microscopy study. Sci Rep 6:
25127
Hossain MK, Strezov V, Yin Chan K, Nelson PF (2010) Agronomic properties of wastewater
sludge biochar and bioavailability of metals in production of cherry tomato (Lycopersicon
esculentum). Chemosphere 78:1167–1171
Hossain MK, Strezov Vladimir V, Chan KY, Ziolkowski A, Nelson PF (2011) Influence of
pyrolysis temperature on production and nutrient properties of wastewater sludge biochar. J
Environ Manag 92:223–228
Houben D, Evrard L, Sonnet P (2013) Mobility, bioavailability and pH-dependent leaching of
cadmium, zinc and lead in a contaminated soil amended with biochar. Chemosphere 92:1450–
1457
Huang R, Zhang B, Saad EM, Ingall ED, Tang Y (2018) Speciation evolution of zinc and copper
during pyrolysis and hydrothermal carbonization treatments of sewage sludges. Water Res 132:
260–269
Inyang MI, Gao B, Yao Y, Xue Y, Zimmerman A, Mosa A, Pullammanappallil P, Ok YS, Cao X
(2016) A review of biochar as a low-cost adsorbent for aqueous heavy metal removal. Crit Rev
Environ Sci Technol 4:406–433
Iturbe R, López J (2015) Bioremediation for a soil contaminated with hydrocarbons. J Pet Environ
Biotechnol 06
Iwasaki T, Suzuki S, Kojima T (2014) Influence of biomass pyrolysis temperature, heating rate and
type of biomass on produced char in a fluidized bed reactor. Energy Environ Res 4:64–72
Jin H, Capareda S, Chang Z, Gao J, Xu Y, Zhang J (2014) Biochar pyrolytically produced from
municipal solid wastes for aqueous As(V) removal: adsorption property and its improvement
with KOH activation. Bioresour Technol 169:622–629
Jung C, Park J, Hun K, Park S, Heo J, Her N, Oh J, Yun S, Yoon Y (2013) Adsorption of selected
endocrine disrupting compounds and pharmaceuticals on activated biochars. J Hazard Mater
263:702–710
Kambo HS, Dutta A (2015) A comparative review of biochar and hydrochar in terms of production,
physico-chemical properties and applications. Renew Sustain Energy Rev 45:359–378
Kang S, Jian-chun J, Dan-dan C (2011) Preparation of activated carbon with highly developed
mesoporous structure from Camellia oleifera shell through water vapor gasification and phos-
phoric acid modification. Biomass Bioenergy 35:3643–3647
Karaj S, Kanyaporn C, Maurer C, Meissner K, Kiatsiriroat T, Müller J (2011) Physico – chemical
characterization of Biochar products from Jatropha curcas L. shells, press cake and solid biogas
digestate. In: Asia Pacific Biochar conference, APBC Kyoto 2011
Kargar M, Clark OG, Hendershot WH, Jutras P, Prasher SO (2015) Immobilization of trace metals
in contaminated urban soil amended with compost and biochar. Water Air Soil Pollut 226:191
Khan S, Wang N, Reid BJ, Freddo A, Cai C (2013) Reduced bioaccumulation of PAHs by Lactuca
satuva L. grown in contaminated soil amended with sewage sludge and sewage sludge derived
biochar. Environ Pollut 175:64–68
Kim WK, Shim T, Kim YS, Hyun S, Ryu C, Park YK, Jung J (2013) Characterization of cadmium
removal from aqueous solution by biochar produced from a giant Miscanthus at different
pyrolytic temperatures. Bioresour Technol 138:266–270
Kumpiene J (2010) Trace element immobilization in soil using amendments
Kuppusamy S, Thavamani P, Megharaj M, Venkateswarlu K, Naidu R (2016) Agronomic and
remedial benefits and risks of applying biochar to soil: current knowledge and future research
directions. Environ Int 87:1–12
Kuzyakov Y, Subbotina I, Chen H, Bogomolova I, Xu X (2009) Black carbon decomposition and
incorporation into soil microbial biomass estimated by 14C labeling. Soil Biol Biochem 41:210–
219
290 S. Wongrod et al.
Laghari M, Naidu R, Xiao B, Hu Z, Mirjat MS, Hu M, Kandhro MN, Chen Z, Guo D, Jogi Q, Abudi
ZN, Fazal S (2016) Recent developments in biochar as an effective tool for agricultural soil
management: a review. J Sci Food Agric 96:4840–4849
Lehmann J, Joseph S (2009) Biochar for environmental management: an introduction. Sci Technol
1:1–12
Liu T, Liu B, Zhang W (2014) Nutrients and heavy metals in biochar produced by sewage sludge
pyrolysis: its application in soil amendment. Polish J Environ Stud 23:271–275
Liu CH, Chuang YH, Li H, Boyd SA, Teppen BJ, Gonzalez JM, Johnston CT, Lehmann J, Zhang
W (2019) Long-term sorption of lincomycin to biochars: the intertwined roles of pore diffusion
and dissolved organic carbon. Water Res 161:108–118
Malińska K, Golańska M, Cáceres R, Rorat A, Weisser P, Ewelina Ś (2017) Biochar amendment for
integrated composting and vermicomposting of sewage sludge—the effect of biochar on the
activity of Eisenia fetida and the obtained vermicompost. Bioresour Technol 225:206–214
Méndez A, Gómez A, Paz-Ferreiro J, Gascó G (2012) Effects of sewage sludge biochar on plant
metal availability after application to a Mediterranean soil. Chemosphere 89:1354–1359
Meng J, Wang L, Liu X, Wu J, Brookes PC, Xu J (2013) Physicochemical properties of biochar
produced from aerobically composted swine manure and its potential use as an environmental
amendment. Bioresour Technol 142:641–646
Mohan D, Sarswat A, Ok YS, Pittman CU (2014) Organic and inorganic contaminants removal
from water with biochar, a renewable, low cost and sustainable adsorbent—a critical review.
Bioresour Technol 160:191–202
Monlau F, Sambusiti C, Antoniou N, Barakat A, Zabaniotou A (2015) A new concept for enhancing
energy recovery from agricultural residues by coupling anaerobic digestion and pyrolysis
process. Appl Energy 148:32–38
Mukherjee S, Weihermuller L, Tappe W, Hofmann D, Koppchen S, Laabs V, Vereecken H,
Burauel P (2016) Sorption-desorption behaviour of bentazone, boscalid and pyrimethanil in
biochar and digestate based soil mixtures for biopurification systems. Sci Total Environ 559:63–
73
Nartey OD, Zhao B (2014) Biochar preparation, characterization, and adsorptive capacity and its
effect on bioavailability of contaminants: an overview. Adv Mater Sci Eng 2014:1–12
Neumann J, Binder S, Apfelbacher A, Gasson JR, Ramírez García P, Hornung A (2014) Production
and characterization of a new quality pyrolysis oil, char and syngas from digestate – Introducing
the thermo-catalytic reforming process. J Anal Appl Pyrolysis 113:137–142
Novotny EH, Branco de Freitas Maia CM, de Melo Carvalho MT, Emöke Madari B (2015)
Biochar: pyrogenic carbon for agricultural use—a critical review. Rev Bras Ciência do Solo
39:321–344
Oh SY, Seo YD (2016) Sorption of halogenated phenols and pharmaceuticals to biochar: affecting
factors and mechanisms. Environ Sci Pollut Res 23:951–961
Ok YS, Uchimiya SM, Chang SX, Bolan N (2015) Biochar: production, characterization, and
applications. CRC, Boca Raton
Paz-Ferreiro J, Liang C, Fu S, Mendez A, Gasco G (2015) The effect of biochar and its interaction
with the earthworm Pontoscolex corethrurus on soil microbial community structure in tropical
soils. PLoS One 10:1–11
Peng W, Pivato A (2017) Sustainable management of digestate from the organic fraction of
municipal solid waste and food waste under the concepts of back to earth alternatives and
circular economy. Waste Biomass Valorization 10:465–481
Pituello C, Francioso O, Simonetti G, Pisi A, Torreggiani A, Berti A, Morari F (2014) Character-
ization of chemical–physical, structural and morphological properties of biochars from
biowastes produced at different temperatures. J. Soils Sediment 15:792–804
Qiu Y, Zheng Z, Zhou Z, Sheng GD (2009) Effectiveness and mechanisms of dye adsorption on a
straw-based biochar. Bioresour Technol 100:5348–5351
10 Biochar Produced from Organic Waste Digestate and Its Potential. . . 291
Rajapaksha AU, Vithanage M, Zhang M, Ahmad M, Mohan D, Chang SX, Ok YS (2014) Pyrolysis
condition affected sulfamethazine sorption by tea waste biochars. Bioresour Technol 166:303–
308
Rajapaksha AU, Chen SS, Tsang DCW, Zhang M, Vithanage M, Mandal S, Gao B, Bolan NS, Ok
YS (2016) Engineered/designer biochar for contaminant removal/immobilization from soil and
water: potential and implication of biochar modification. Chemosphere 148:276–291
Rehrah D, Bansode RR, Hassan O, Ahmedna M (2015) Physico-chemical characterization of
biochars from solid municipal waste for use in soil amendment. J Anal Appl Pyrolysis 118:
42–53
Rodriguez M (2010) Biochar as a strategy for sustainable land management, poverty reduction and
climate change mitigation/adaptation? Carbon NY
Singh B, Singh BP, Cowie AL (2010) Characterisation and evaluation of biochars for their
application as a soil amendment. Aust J Soil Res 48:516–525
Singh BP, Cowie AL, Smernik RJ (2012) Biochar carbon stability in a clayey soil as a function of
feedstock and pyrolysis temperature. Environ Sci Technol 46:11770–11778
Sizmur T, Quilliam R, Puga AP, Moreno-Jiménez E, Beesley L, Gomez-Eyles JL (2015) Applica-
tion of biochar for soil remediation. Soil Sci Soc Am Spec Publ 63:1–30
Sohi SP, Krull E, Lopez-Capel E, Bol R (2010) A review of biochar and its use and function in soil.
Adv Agron 105:47–82
Spokas KA (2010) Review of the stability of biochar in soils: predictability of O:C molar ratios.
Carbon Manag 1:289–303
Stavropoulos GG, Samaras P, Sakellaropoulos GP (2008) Effect of activated carbons modification
on porosity, surface structure and phenol adsorption. J Hazard Mater 151:414–421
Stefaniuk M, Oleszczuk P, Różyło K (2017) Co-application of sewage sludge with biochar
increases disappearance of polycyclic aromatic hydrocarbons from fertilized soil in long term
field experiment. Sci Total Environ 599–600:854–862
Tampio E, Salo T, Rintala J (2016) Agronomic characteristics of five different urban waste
digestates. J Environ Manag 169:293–302
Tan X, Liu Y, Zeng G, Wang X, Hu X, Gu Y, Yang Z (2015) Application of biochar for the removal
of pollutants from aqueous solutions. Chemosphere 125:70–85
Tsai WT, Liu SC, Chen HR, Chang YM, Tsai YL (2012) Textural and chemical properties of swine-
manure-derived biochar pertinent to its potential use as a soil amendment. Chemosphere 89:
198–203
van Zwieten L, Kimber S, Morris S, Chan KY, Downie A, Rust J, Joseph S, Cowie A (2010) Effects
of biochar from slow pyrolysis of papermill waste on agronomic performance and soil fertility.
Plant Soil 327:235–246
Verheijen F, Jeffery S, Bastos AC, Van Der Velde M, Diafas I (2010) Biochar application to soils.
JRC scientific and technical report
Vithanage M, Rajapaksha AU, Tang X, Thiele-Bruhn S, Kim KH, Lee S-E, Ok YS (2014) Sorption
and transport of sulfamethazine in agricultural soils amended with invasive-plant-derived
biochar. J Environ Manag 141:95–103
Wang Z, Han L, Sun K, Jin J, Ro KS, Libra JA, Liu X, Xing B (2016) Sorption of four hydrophobic
organic contaminants by biochars derived from maize straw, wood dust and swine manure at
different pyrolytic temperatures. Chemosphere 144:285–291
Waqas M, Khan S, Qing H, Reid BJ, Chao C (2014) The effects of sewage sludge and sewage
sludge biochar on PAHs and potentially toxic element bioaccumulation in Cucumis sativa
L. Chemosphere 105:53–61
Weyers SL, Spokas K a (2011) Impact of biochar on earthworm populations: a review. Appl
Environ Soil Sci 2011:1–12
292 S. Wongrod et al.
Williams M, Martin S, Kookana RS (2015) Sorption and plant uptake of pharmaceuticals from an
artificially contaminated soil amended with biochars. Plant Soil 395:75–86
Wongrod S, Simon S, Guibaud G, Lens PNL, Pechaud Y, Huguenot D, van Hullebusch ED (2018a)
Lead sorption by biochar produced from digestates: consequences of chemical modification and
washing. J Environ Manag 219:277–284
Wongrod S, Simon S, van Hullebusch ED, Lens PNL, Guibaud G (2018b) Changes of sewage
sludge digestate-derived biochar properties after chemical treatments and influence on As(III
and V) and Cd(II) sorption. Int Biodeter Biodegr 135:96–102
Wuana RA, Okieimen FE (2011) Heavy metals in contaminated soils: a review of sources,
chemistry, risks and best available strategies for remediation. ISRN Ecol 2011:1–20
Xu RK, Xiao SC, Yuan JH, Zhao AZ (2011) Adsorption of methyl violet from aqueous solutions by
the biochars derived from crop residues. Bioresour Technol 102:10293–10298
Xu X, Hu X, Ding Z, Chen Y, Gao B (2017) Waste-art-paper biochar as an effective sorbent for
recovery of aqueous Pb(II) into value-added PbO nanoparticles. Chem Eng J 308:863–871
Xue Y, Gao B, Yao Y, Inyang M, Zhang M, Zimmerman AR, Ro KS (2012) Hydrogen peroxide
modification enhances the ability of biochar (hydrochar) produced from hydrothermal carbon-
ization of peanut hull to remove aqueous heavy metals: batch and column tests. Chem Eng J
200–202:673–680
Yang R, Liu G, Xu X, Li M, Zhang J, Hao X (2011) Surface texture, chemistry and adsorption
properties of acid blue 9 of hemp (Cannabis sativa L.) bast-based activated carbon fibers
prepared by phosphoric acid activation. Biomass Bioenergy 35:437–445
Yang Y, Zhang W, Qiu H, Tsang DCW, Morel JL, Qiu R (2016) Effect of coexisting Al(III) ions on
Pb(II) sorption on biochars: role of pH buffer and competition. Chemosphere 161:438–445
Yuan H, Lu T, Huang H, Zhao D, Kobayashi N, Chen Y (2015) Influence of pyrolysis temperature
on physical and chemical properties of biochar made from sewage sludge. J Anal Appl Pyrolysis
112:284–289
Zhang X, Wang H, He L, Lu K, Sarmah A, Li J, Bolan NS, Pei J, Huang H (2013) Using biochar for
remediation of soils contaminated with heavy metals and organic pollutants. Environ Sci Pollut
Res 20:8472–8483
Zhang QZ, Dijkstra FA, Liu XR, Wang YD, Huang J, Lu N (2014) Effects of biochar on soil
microbial biomass after four years of consecutive application in the north China Plain. PLoS
One 9:1–8
Zhang X, Sarmah AK, Bolan NS, He L, Lin X, Che L, Tang C, Wang H (2015) Effect of aging
process on adsorption of diethyl phthalate in soils amended with bamboo biochar. Chemosphere
142:28–34
Zhang W, Huang X, Jia Y, Rees F, Tsang DCW, Qiu R, Wang H (2016) Metal immobilization by
sludge-derived biochar: roles of mineral oxides and carbonized organic compartment. Environ
Geochem Health 39(2):379–389
Zheng H, Wang Z, Zhao J, Herbert S, Xing B (2013) Sorption of antibiotic sulfamethoxazole varies
with biochars produced at different temperatures. Environ Pollut 181:60–67
Zielińska A, Oleszczuk P (2015) The conversion of sewage sludge into biochar reduces polycyclic
aromatic hydrocarbon content and ecotoxicity but increases trace metal content. Biomass
Bioenergy 75:235–244
Zielińska A, Oleszczuk P (2016) Attenuation of phenanthrene and pyrene adsorption by sewage
sludge-derived biochar in biochar-amended soils. Environ Sci Pollut Res 23:21822–21832
Zielińska A, Oleszczuk P, Charmas B, Skubiszewska-Zięba J, Pasieczna-Patkowska S (2015)
Effect of sewage sludge properties on the biochar characteristic. J Anal Appl Pyrolysis
112:201–213
Part V
Integration of AD in Biorefineries
Chapter 11
Integration of Bio-electrochemical Systems
with Anaerobic Digestion
Abstract Anaerobic digestion is a well versed technology for treating organic waste
and converting the organic matter into biogas to be subsequently used as fuel or
converted into electricity. Over the years, anaerobic digestion has been widely used
for the treatment of industrial and domestic wastewaters as well as organic solid
waste, while simultaneously generating biogas. On the other hand,
bio-electrochemical systems (BES) are one of the prominent and upcoming ways
of harnessing organic matter into bioelectricity and industrially valuable products,
such as hydrogen, methane, biofuels, and hydrogen peroxide. This chapter reviews
the hybridization of BES and anaerobic digestion for enhanced pollutant removal
and pollutant monitoring in the form of nutrient removal and recovery, polishing of
anaerobically treated effluent, biogas upgrading, application of biosensors, and
merger of BES in anaerobic digestion for emerging contaminant removal. Addition-
ally, a brief discussion on the future prospect of the synergy of these two technol-
ogies has also been highlighted.
M. M. Ghangrekar (*)
Department of Civil Engineering, Indian Institute of Technology Kharagpur, Kharagpur, West
Bengal, India
School of Environmental Science and Engineering, Indian Institute of Technology Kharagpur,
Kharagpur, West Bengal, India
e-mail: [email protected]
S. M. Sathe
Department of Civil Engineering, Indian Institute of Technology Kharagpur, Kharagpur, West
Bengal, India
C. N. Khuman
School of Environmental Science and Engineering, Indian Institute of Technology Kharagpur,
Kharagpur, West Bengal, India
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 295
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_11
296 M. M. Ghangrekar et al.
11.1 Introduction
The urge of finding alternative and sustainable energy sources to replace ever-
depleting fossil fuels is a major global issue for sustainable development. The use
of fossil fuels is not environmentally friendly due to the emission of harmful gases
into the atmosphere. Therefore, the use of renewable energy sources such as solar,
wind, geothermal, and tidal are becoming increasingly popular now-a-days (Jingura
and Matengaifa 2009). Organic waste or biomass is also one of the promising
renewable energy sources, which is considered as an environmentally friendly
approach (Berndes et al. 2003; Lin et al. 2018; Stephen and Periyasamy 2018). It
is estimated that 1 kg of carbohydrates represents 1.06 kg of chemical oxygen
demand (COD), equivalent to 4.41 kWh of energy (Rabaey and Verstraete 2005).
Anaerobic digestion (AD) is one of the efficient and proven methods for the
treatment of organic matter present in solid waste and wastewater. It is specifically
preferred for the treatment of high strength waste in terms of COD due to its low
sludge generation, low energy requirement, and formation of valuable methane as
end product (Chen et al. 2008; Kiyasudeen et al. 2016). The AD of organic matter
takes place via hydrolysis, acidogenesis, acetogenesis, and methanogenesis,
resulting in the generation of biogas, which comprises of CH4 and CO2 and other
gases such as H2S, H2 and N2 (Gujer and Zehnder 1983; Zhang et al. 2014a).
Methane generated from the process can be either used as fuel or converted into
electricity. AD is a complex process involving a plethora of microorganisms, in
which the end-products of the initial bacterial reaction are often the substrate for
subsequent bioconversion. Hence, a balance between the microbial population and
the substrate concentration is vital for the stable operation of AD. In the initial three
stages (hydrolysis, acidogenesis, and acetogenesis), the organic matter is broken
down into smaller molecules via the formation of long chain volatile fatty acids
(VFAs) to finally short chain VFA and acetic acid; whereas, in the final stage
(methanogenesis), methane is produced by the methanogens (acetoclastic and
hydrogenotropic) (Gavala et al. 2003; Gude 2018).
The process of AD is vastly affected by variation in temperature, substrate pH,
hydraulic retention time (HRT), organic loading rate (OLR), and C/N ratio (Mao
et al. 2015). Apart from this, the failure of AD is often linked to toxic effects induced
by inhibitory substances. The presence of ammonia, sulfur, heavy metals, other ions
(Na+, K+, Mg2+, Ca2+, Al3+), and organic compounds (such as alkyl benzenes,
chlorophenols, surfactants, and alkanes) beyond the threshold limit can cause an
imbalance in the AD (Chen et al. 2008). Any imbalance in the process can be related
to reduced CH4 generation and deteriorated organic matter removal efficiency (Chen
et al. 2008). This necessitates the requirement of pre-treatments to make the waste fit
for treatment in AD.
Bio-electrochemical systems (BES) are a new generation of wastewater treatment
techniques capable of recovery of valuables and/or bioenergy production. In the
anodic and cathodic chambers of a BES, respectively oxidation and reduction takes
place. The organic matter from the waste is oxidized in the anodic chamber using
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 297
During AD, microorganisms decompose the organic matter into simpler molecules
with the release of energy in the form of electrons and protons. This energy can be
harvested using a novel technology known as BES. Compared to AD, the BES is a
recent technology that is capable of converting dissolved organic matter from
wastewater into bioelectricity or bioenergy. Though the first instance of electricity
generation using microorganisms was demonstrated by Potter in 1911 (Potter 1911),
the subsequent in depth research and widespread applications were initiated only in
the late twentieth century. Owing to the capability of one-step bioelectricity produc-
tion, BES become an attractive option for the conversion of biomass/organic matter
to electrical energy or by applying small external potential other value-added
products can be recovered.
A BES is an electrochemical system catalysed by the action of microorganisms
for the treatment of polluted soil and wastewater, while harvesting bioenergy or
synthesizing value-added products such as biofuels, H2O2, H2 and CH4. Addition-
ally, they have also been applied for the sequestration of CO2 with the concurrent
production of biochemical and biofuels in microbial electrosynthesis cells (Liu et al.
2005; Jeremiasse et al. 2010; Das et al. 2020). Majorly, BES may be galvanic as in
the case of microbial fuel cells (MFCs), which allow the conversion of chemical
energy of organic matter into electricity or electrolytic as in the case of microbial
electrolytic cells (MECs), where electricity is externally supplied to the system to
facilitate recovery of industrially valuable products (biohydrogen, biomethane, and
hydrogen peroxide).
In a nutshell, BES consists of a single pair of electrodes or multiple pairs of
electrodes, an ion exchange membrane (may not be necessary for a membrane-less
BES configuration), an external load (in case of MFC), or an external power supply
(in case of MEC), and microorganisms especially electrogens (Logan et al. 2006).
The electrogens are capable of electron transfer to the electrodes from their cell
(known as exoelectrogens) and from the electrodes into their cell (known as
endoelectrogens or electrotrophs) (Semenec and Franks 2015; Kang et al. 2021).
298 M. M. Ghangrekar et al.
In a typical double chamber MFC, the anode is dipped inside wastewater, while the
cathode is placed in air-saturated water (aqueous cathode) or exposed to air (air
cathode). The anode and the cathode are connected to an external load. The anolyte
(wastewater) and the catholyte (air-saturated water or air) are separated by a proton
exchange membrane (PEM) or cation exchange membrane (Fig. 11.1a). Oxidation of
the organic matter occurs in the anodic chamber under anaerobic conditions using
microorganisms, thereby releasing electrons, protons, and CO2. The presence of
terminal electron acceptors (generally oxygen) in the catholyte attracts the electrons
towards the cathodic compartment. However, the PEM allows only the passage of
protons and prohibits the flow of electrons and organic matter towards the cathode.
The electrons flow towards the cathode through the external circuit and create a
charge imbalance between the anolyte and the catholyte. To maintain charge equi-
librium the protons flow through the PEM towards the cathode to combine with O2
and electrons forming H2O or H2O2 (Eqs. 11.1 and 11.2). Thus, electricity is
harvested due to the flow of electrons through the external circuit (Kim et al.
2004; Liu et al. 2005).
Power supply
e− External Load e− e− e−
a b
−
Substrate e− Substrate e e−
Separator
e−
Separator
O2
H2
Anode
Anode
H2O
Cathode
Cathode
H+ H+ H+ H+
CO2 CO2
Microorganisms Microorganisms
Anodic chamber Cathodic chamber Anodic chamber Cathodic chamber
Fig. 11.1 (a) Typical dual-chambered microbial fuel cell and (b) dual-chambered microbial
electrolysis cell
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 299
In contrast to MFC, electrical power is externally supplied into the electrolytic BES
(Fig. 11.1b). At the anode, oxidation of organic matter catalysed by microorganisms
and water hydrolysis takes place. The electrons flow toward the cathode through the
external power circuit, while the protons migrate to the cathode through the PEM,
similar to that of MFC. The MECs are capable of producing industrial valuable
products, such as H2, H2O2 and organic acids, in cathodic chamber along with
simultaneous removal of organic matter from wastewater in anodic chamber
(Jeremiasse et al. 2010; Jafary et al. 2015). However, the energy required for
bioconversion of organic matter into industrial valuable products is lesser in MEC
than what is required in conventional electrochemical methods due to the availability
of electrons and protons released during microbial decomposition of organic matter
(Zhang et al. 2014b; Kuntke et al. 2018).
BES have been widely explored for the removal and recovery of nutrients (nitrogen,
phosphorus, sulfur) from the wastewater that cannot be completely removed using
AD. These nutrients have agricultural use and since the recovery of nutrients is
possible from the waste source; the BES perfectly fits into the concept of waste to
wealth.
The presence of nutrients and common elements is vital for AD and depending on
the particular concentration it can benefit or harm the digestion process. In some of
the high strength waste streams in terms of COD, such as concentrate from waste-
water sludge, food waste, and leachate, the concentration of total ammonia nitrogen
is often considerably higher than that detected in the domestic wastewater. It was
observed that a high concentration of total ammonia causes excessive inhibition of
methanogenesis (Sheng et al. 2013). The digestion process in the anaerobic digester
is most commonly disturbed by the sudden high ammoniacal nitrogen concentration
in the influent waste. High ammonia concentrations in the influent cause a consid-
erable increase in the VFA production resulting into reduced methane production
(Borja et al. 1996). With the rapidly increasing VFA, the digester might not have
300 M. M. Ghangrekar et al.
enough buffering capacity to maintain the pH in the working range of 6.8–7.2 that is
required for the methanogens, and can thus lead to the failure of AD. In such cases
(high ammoniacal nitrogen in the influent), instead of removal of ammonia in the
form of N2 gas, its recovery (as fertilizer) can be a beneficial option, considering the
valuable applications of the recovered products (ammonium sulfate, ammonium
bicarbonate) in the agricultural sector.
For such wastes, the application of a specific pre-treatment technique prior to AD,
which is targeted for ammonia removal, might be the best option (Kim et al. 2015a).
Over the years, several techniques for the removal of nitrogen from the wastewater
have been developed and implemented, which include nitrification and denitrifica-
tion, anammox process, chemical precipitation, and ion exchange (Peng and Zhu
2006; Zhang et al. 2009b; Lin et al. 2016; Ma et al. 2016). For the wastes that have
high nitrogen content, recovery instead of removal can be a feasible option to utilize
the nitrogen for beneficial uses. In this regard, BES offers an advantage of nitrogen
recovery from the waste at a lower operating cost compared with conventional
systems (Zamora et al. 2017).
In the cathodic chamber of BES, the formation of hydroxyl results in an increase
in pH. The transfer of NH3 and NH4+ through the cation exchange membrane to the
cathodic chamber of BES is due to diffusion and migration, respectively, wherein
increased pH causes conversion of NH4+ to NH3 in the cathodic chamber (Kim et al.
2015b). In this way, NH4+ ions are removed from the anolyte (waste stream). The
formed NH3 can be successively recovered in the form of ammonium sulfate or
ammonium bicarbonate that can be used as a fertilizer. The ammonium sulfate can be
obtained via passing gaseous NH3 through sulfuric acid; whereas the catholyte can
be further concentrated via the use of a tri-chambered system with cation and anion
exchange membranes to obtain solid ammonium bicarbonate (Nancharaiah et al.
2016; Jermakka et al. 2018). The BES have been used for nitrogen removal and its
successive recovery from various wastewater sources, such as urine (Kuntke et al.
2012), landfill leachate (Qin et al. 2016), effluent of sludge dewatering process
(Wu and Modin 2013), and the liquid fraction of pig slurry (Sotres et al. 2015).
The ammonia recovery depends on the imposed electrical potential and it was
observed that under the applied potential (MEC mode), ammonia recovery is higher.
Though the supplemented external energy increases the operation cost, it is still
lower than electrochemical ammonia recovery methods.
In a scaled-up MEC system operated with urine as anolyte (2.6 L of each
catholyte and anolyte), 31% of the total ammonia nitrogen recovery was achieved
under the applied potential of + 0.5 V (against Ag/AgCl reference). The electrical
energy required for the total ammonia nitrogen recovery in the aforementioned
investigation was 4.9 MJ kg N1 (Zamora et al. 2017). On the other hand, consid-
ering only the running electricity and fossil energy requirements, the conventional
biological nitrification and denitrification process requires about 45 MJ kg N1 of
energy in wastewater treatment plants (Maurer et al. 2003). Though complete
removal of the nitrogen fraction is not possible using BES, it can reduce the nitrogen
concentration so that the effluent can be suitable for AD, without negatively affect-
ing the digestion process.
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 301
In one of the investigations, MFC was implemented along with AD for NH4+
removal. In between the recirculation loop of AD, a MFC was installed with Nafion
as a separator. Substrate with 77–124.4 mg of N g1 of total solids was used as a feed
for anaerobic digester. The MFC resulted in a reduction of the NH4+ concentration
from 654 to 436 mg L1. The addition of a MFC caused an increase in methane yield
from 136 to 173 mL CH4 L1 day1. The reduced NH4+ might have improved the
stability of methanogenic consortia leading to an increased methane yield (Inglesby
and Fisher 2012). This shows that integration of BES with AD could reduce the
nutrient load thus it can be a feasible solution to prevent the failure of anaerobic
digestion.
of ammonium nitrogen and phosphate, respectively. This confirmed that the current
generation is important for nutrient removal and higher current due to externally
applied voltage can improve the nutrient removal efficiency in BES (Zhang et al.
2014b).
Many of the industrial wastewaters are the prime source of sulfate. Chemical
precipitation, membrane filtration and biological treatments using sulfate reducing
bacteria are a common way for the removal of sulfate from sulfate laden wastewater.
Although domestic wastewater has a low sulfate concentration, the converted
sulfide remains dissolved in the anaerobically digested effluent that lowers the
overall performance. Hence, a post treatment is also essential to eliminate the sulfide
from the wastewater. Although sulfate in low concentrations does not contribute to
toxicity, a high concentration of sulfate can impart salinization of freshwater bodies
(Lens et al. 1998; Sharma and Kumar 2020). During the AD of wastewater, sulfate is
reduced to sulfide by sulfate reducing bacteria. The generated sulfide can be inhib-
itory to the methane producing bacteria, which reduces the rate of methanogenesis,
and decreases the methane yield by competing with the available carbon and/or
hydrogen (McCartney and Oleszkiewicz 1991). Hence, the sulfate concentration in
the influent wastewater needs to be controlled to prevent the inhibitory effect on
AD. The BES have been considered as among the promising technologies for the
removal of sulfate.
For the removal and recovery of sulfur in MFC, the sulfate reducing bacteria
should work along with electrogens in the anodic chamber. The sulfate present in the
wastewater is reduced to sulfide by sulfate reducing bacteria and it is subsequently
oxidized by sulfide oxidising bacteria to form elemental sulfur (Lee et al. 2014). This
has been demonstrated for the first time by Rabaey et al. (2006) using a MFC while
simultaneously harvesting a power output of 101 mW L1 volume of net anodic
compartment. Granular deposits containing 190 mg of sulfur, corresponding to about
9% of sulfide dose, were observed on the anode. When the same setup of MFC was
coupled with an up-flow anaerobic sludge blanket (UASB) reactor, a total of 98%
sulfide removal was achieved (Rabaey et al. 2006). The coupling of MFC shows that
sulfide can be successfully removed from the anaerobically treated effluent, thus
accomplishing a polishing treatment to the effluent. Another investigation showed
that the performance of a MFC deteriorated when the COD:sulfate ratio was dropped
to less than 0.8. The maximum COD removal efficiency of 79% at a power density of
1136.8 mW m3 was demonstrated using MFC at a COD: sulfate ratio of 0.8
(Ghangrekar et al. 2010).
Chatterjee et al. (2017) used a dual chamber MFC fed with sulfate (200 mg L1)
along with acetate as a carbon source. In the anodic chamber of a MFC, sulfate was
reduced to sulfide, which was observed via an increase in sulfide concentration from
0 to 55 mg L1 at an HRT of 48 h. However, after an additional 24 h, the sulfide
concentration dropped to 3 mg L1. In the absence of oxygen and with reducing
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 303
sulfate concentration, it was concluded that sulfate was reduced to sulfide, which
was subsequently oxidized to elementary sulfur. Energy dispersive X-ray spectros-
copy confirmed the presence of elemental sulfur on the anode with a weight
percentage of 27.45% of the entire surface area (Chatterjee et al. 2017). In the
same investigation, Co3O4 nano-octahedron was used as a cathode catalyst, whereas
a separate MFC without catalyst was operated under the same operating conditions,
to analyse the effect of improved electrical performance and reduced oxygen reduc-
tion reaction overpotential of the sulfide removal. However, it was observed that
except for increased power density and coulombic efficiency, the cathode catalyst
had no significant role in improving the sulfide oxidation (Chatterjee et al. 2017).
In a recent investigation, a tri-chamber MFC comprising of an oxic cathode and
anoxic cathode sharing the same anode was used for nitrogen and sulfate removal.
The system achieved total nitrogen and sulfide removal at rates of 10.0 g N m3 day1
and 206.5 g S m3 day1, respectively (Zhang et al. 2018). The available investi-
gations confirm the efficacy of BES for the removal of sulfate from wastewater,
which can be coupled with AD to aid in the overall treatment process. In the
conventional aerobic biological processes, sulfide is oxidized to sulfate; however,
in the secondary clarifier due to the oxygen depleting environment, the sulfate can
again be reduced to sulfide. On the other hand, using MFC, sulfide is converted to
elemental sulfur, thus offering a better way of addressing sulfide removal (Chatterjee
et al. 2017).
The use of BES in MFC or MEC mode had demonstrated promising results in
terms of removal and recovery of nutrients from the wastewater. The inability of AD
to eliminate nutrients creates need for coupling AD with suitable nutrient removal
techniques. Hence, integration of BES with AD can be a low-cost solution, which
can not only improve the performance of AD, but also be an effective way of nutrient
recovery.
The biogas produced from AD mainly consists of CH4 (55–75%) and CO2
(30–45%). Besides these, other gases such as H2S, N2, H2 along with H2O (moisture)
can also be present in the biogas (Igoni et al. 2008). However, CH4 is the only
component of interest in biogas as a source of fuel. The presence of CO2, H2S, and
H2O leads to a lower calorific value of biogas. Hence, biogas needs to be processed
before utilising it as a fuel source. Biogas processing is done in two broad processes,
viz. biogas cleaning (removal of impurities such as H2O and H2S) and biogas
upgrading (removal or transformation of CO2 into CH4 to increase the CH4 content
in the biogas) (Angelidaki et al. 2018; Fu et al. 2020).
Biogas upgrading is more concerning than cleaning as CO2 contributes a large
percentage of the total biogas generated. Biogas can be upgraded through physical
absorption, chemical absorption, cryogenic separation, and membrane separation
(Miltner et al. 2017; Rafiee et al. 2021). Biological methods of biogas upgrading
304 M. M. Ghangrekar et al.
the cathode. Further, the transcriptomic analysis revealed that Methanothrix uses the
CO2 reduction pathway that stimulates direct electron transfer in MEC for enhancing
the CH4 content of biogas (Liu et al. 2019). These investigations showed that BES
presents a potentially sustainable means for in-situ or ex-situ biogas upgrading.
The AD as a sole unit cannot be used as a stand-alone treatment for wastewater. For
high strength waste in terms of COD, the effluent of AD still contains a considerable
amount of residual organic matter, thus requiring the need for a second stage of
treatment. BES can be used in such cases as a follow-up treatment for organic matter
removal and energy recovery. On the other hand, for low strength wastewater in
terms of COD, the effluent contains nutrients and pathogens; thus, BES can be used
as an add-on secondary treatment unit or as a polishing treatment unit or both for the
effluent of AD.
Zhang et al. (2012a) reported treatment of organic matter and sulfate in an UASB
reactor and MFC. The combined system of UASB followed by MFC achieved 69.9
and 81.8% of COD and sulfate removal, respectively. The MFC harvested a max-
imum power density of 989.6 mW m2 at a COD/sulfate ratio of 2.3 and HRT of
54.3 h (Zhang et al. 2012a). In another investigation, four MFCs in series were used
as a polishing treatment for the biohydrogen and biomethanation process. The MFCs
were operated with varied OLR of 0.036–6.149 g COD L1 day1, resulting in a
COD removal efficiency of 35.1–4.4%. The highest coulombic efficiency of 60%
was attained at an OLR of 0.572 g COD L1 day1 simultaneously producing a
maximum power density of 3.1 W m3 (Fradler et al. 2014).
In another combination, an UASB-MFC-biological aerated filter was developed
for the molasses wastewater treatment. The UASB reactor was used for COD
removal and the MFC for generating bioelectricity simultaneously degrading COD
as well as oxidation of the generated sulfide; whereas the colour removal and phenol
degradation was accomplished in the biological aerated filter. The overall removal
efficiency of 53.2, 52.7, and 41.1% for COD, sulfate, and colour, respectively, was
achieved in the combined system. The MFC in this combined system produced a
maximum power density of 1410.2 mW m2 (Zhang et al. 2009a). Various combi-
nations have been tried to integrate BES as downstream treatment to AD in order to
meet the discharge norms for the disposal of the treated effluent. Additionally, the
harvested bioelectricity from the MFCs can be used for on-site applications. The
follow-up treatment in BES also offers an advantage for not only the removal of
carbonaceous organic matter but also nutrient removal and recovery as explained in
Sect. 11.3.1.
306 M. M. Ghangrekar et al.
Even after secondary biological treatment targeted for the removal of carbonaceous
organic matter and nutrients from the wastewater, pathogens and emerging contam-
inants (ECs) persist in the effluent. The BES can be implemented to eliminate these
pollutants from the secondary or tertiary treated effluents. The role of downstream
BES-based technology is of paramount importance because AD is incapable of
eliminating the ECs, which could result in discharge of ECs into the receiving
water bodies. Pharmaceuticals, dyes, personal care products, estrogens, aromatic
hydrocarbons, surfactants, and pesticides are amongst the ECs that have been
successfully removed using BES (Chakraborty et al. 2020; Wu et al. 2020). Three
different possible approaches for degrading the ECs from the wastewater using
BES are: (1) anodic degradation; (2) cathodic degradation; and (3) anodic followed
by cathodic degradation.
The ECs can be used as a co-substrate by the anodic microbial consortia along
with organic matter. However, for the low concentration of ECs (generally detected
in the domestic wastewater), cathodic degradation is considered as an attractive
option (Fig. 11.2); wherein, the degradation is achieved using in-situ generated H2O2
or •OH. In an air-cathode MFC when penicillin (50 mg L1) was used with glucose
(1000 mg L1) as a feed, a maximum power density of 101.2 W m3 was generated.
The MFC achieved 98% penicillin removal in 24 h of operation. The maximum
power generated was 6.8-fold higher than that of MFC operated using solely glucose
(1000 mg L1) as a feed. This might be due to the direct electron transfer from the
bacterial cell to the anode by increasing the electron permeability of the bacterial cell
membrane (Wen et al. 2011b). Similarly, the addition of ceftriaxone and Tween-80
in the anodic chamber caused an increase in the electrical performance of MFC,
Recirculation of anolyte
to cathodic chamber
MFC/MEC
e−
Outlet
CO2 C
A A H2O2
Anaerobic N
O
T
H
O 2e− Treated
digester D
E D
E
effluent
OM O2
H+
Inlet
OM: organic matter
Fig. 11.2 Bio-electrochemical system as post treatment to anaerobic digester (second stage
biological treatment followed by tertiary treatment in cathodic chamber)
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 307
simultaneously achieving its removal (Wen et al. 2011a, c). However, not all ECs
showed a positive correlation with power generation. The addition of metronidazole
(10 mg L1) as feed along with glucose caused a reduction in the maximum power
density from 141.94 to 99.23 mW m2 in a dual chamber MFC. However, the MFC
demonstrated 85.4% of metronidazole removal in 24 h (Song et al. 2013).
Due to complexity, the ECs require advanced treatment processes for their
elimination from the wastewater. In the cathodic chamber of BES, the advanced
oxidation processes can be initiated for the removal of ECs. In BES, the Fenton
reaction can be carried via in-situ H2O2 synthesis supplemented with the use of ferric
ions. The first instance of the cathodic Fenton reaction in BES was undertaken by
Zhu and Ni (2009) for the degradation of p-nitrophenol using scrap iron as Fe source.
At a retention time of 12 h, 85% of TOC was removed simultaneously generating
143 mW m2 of maximum power density (Zhu and Ni 2009). In another investiga-
tion, sulfamethoxazole was removed in the anodic chamber followed by cathodic
treatment in a MFC. The combined process demonstrated 94.66% of sulfamethox-
azole removal in 48 h, in which 56.28% removal was in the anodic chamber and
remaining in the cathodic chamber due to Fenton oxidation achieved from carbon
nanotubes, stainless steel mesh, and γ-FeOOH cathode (Li et al. 2020). The bio-
electro-Fenton has been used for the degradation of dyes, antibiotics, pharmaceuti-
cals, and estrogens (Kahoush et al. 2018). It can be used as a polishing treatment to
effluents from an anaerobic digester for the removal of ECs, which cannot be
removed by the AD.
The BES has an edge over conventional AD for the removal of ECs due to the
presence of the anode, which acts as an inexhaustible electron sink. Therefore, the
merger of BES in an anaerobic digester might be one of the options for bioremedi-
ation of recalcitrant contaminants from wastewater. In experiments conducted by Liu
et al. (2011b), a zero valent iron bed with graphite plate electrodes was embedded in
an anaerobic reactor. With the increase in applied voltage, the COD and colour
removal was improved. The bacteriological analysis confirmed the occurrence of azo
dye degrading microbes in the anaerobic reactor (Liu et al. 2011b). Similar obser-
vations of improved COD and azo dye removal were reported in an anaerobic reactor
packed with a pair of Fe-graphite plate electrodes. With 83.4% of COD and 84.7%
dye removal, the Fe-graphite electrode embedded anaerobic reactor outperformed
the sole anaerobic reactor, in which the COD and dye removal efficiency was 76.4
and 50.7%, respectively (Zhang et al. 2012b).
More recently, Shen et al. (2014) tried integration of BES in a UASB reactor for
removal of p-nitrophenol in which a graphite felt cathode and anode were horizon-
tally placed inside the UASB reactor. The results demonstrated that the p-
nitrophenol removal rate in the BES-UASB setup (6.77 M m3 day1) was higher
compared to the UASB reactor only. Also the p-aminophenol (final degradation
308 M. M. Ghangrekar et al.
Apart from the removal of ECs, the cathodic chamber of BES has been used for
disinfection as well. In an investigation conducted by Jadhav et al. (2014), anolyte
was circulated in the cathodic chamber of the MFC; wherein sodium hypochlorite in
different dosage (0.67–3 g L1) was used as catholyte to facilitate disinfection.
Irrespective of the sodium hypochlorite dose, about 99.5% of disinfection was
achieved with the initial bacterial count of 2.8 106 colony-forming units per
100 mL of wastewater. The addition of sodium hypochlorite in 3 g L1 dose
achieved a 85% increase in sustainable power density (80 mW m2) compared to
the 0.67 g L1 dose of sodium hypochlorite. This states that the addition of
chemicals (such as disinfectants) with high redox potential in the catholyte not
only achieved disinfection but also increased the power output of MFC (Jadhav
et al. 2014).
Instead of externally adding the disinfectant, the cathodic disinfection can be
achieved in BES via in-situ synthesis of H2O2 at the cathode. Using oxygen as a
terminal electron acceptor, the cathodic reaction can be progressed via 2e or 4e
based on the cathode material and/or cathode catalyst. Although the 4e pathway is
preferred in BES in terms of electrical performance, the 2e pathway has the
advantage of in-situ H2O2 synthesis at the cathode. The treated anolyte of BES
can be circulated into the cathodic chamber for disinfection or catholyte (high
concentration of H2O2) and treated anolyte can be transferred to a separate offline
chamber for disinfection. Of these, the second option of offline disinfection is better,
as direct circulation of the anolyte into the cathodic chamber can lead to biofouling
on the cathode. The biofouling on the cathode reduces the availability of dissolved
oxygen for the cathodic reaction, thereby increasing the overpotential losses and
reducing the electrical performance (Noori et al. 2019). Nevertheless, while targeting
the two electron pathway of ORR the power produced from the MFC will reduce due
to a lower redox potential of this ORR, however it will offer the advantage of
improved wastewater treatment.
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 309
For the in-situ cathodic H2O2 synthesis different carbon-based cathodes such as
carbon felt (Liu et al. 2011a), graphite plate (Li et al. 2017), graphite (Sim et al.
2015), nitrogen-doped graphitic carbon (Asghar et al. 2017) have been implemented.
Additionally, different metal-based and metallic nanoparticles ranging from Pd/C
(Wang and Wang 2007), Au-Pd (Edwards and Hutchings 2008), and Ni-Pd (Gupta
et al. 2020) have also been successful for cathodic synthesis of H2O2 in BES. The
H2O2 production can be achieved in MFC as well as MEC. However, MEC has an
edge over MFC in terms of the rate of H2O2 generation (Chung et al. 2020).
Over the years, several researchers have tried altering the operating conditions,
such as pH, catholyte, anolyte, imposed potential and cathode catalysts in BES to
maximize the yield of H2O2 in the cathodic chamber (Fu et al. 2010; Sim et al. 2015;
Gupta et al. 2020). Arends et al. (2014) developed a wastewater treatment system
comprising of a wetland with a BES, capable of synthesizing cathodic H2O2. It was
observed that after 1 h of contact time, the produced H2O2 of 0.1% strength was able
to reduce the total coliforms to less than <75 CFU per mL (Arends et al. 2014). In
one of the investigations, Zhou et al. (2018) used bio-electro-Fenton for disinfection
of wastewater in the cathodic chamber and monitoring was performed in terms of
E. coli inactivation. The disinfection capability was optimized by varying the applied
voltage, iron concentration, and catholyte pH. Under acidic conditions (pH of 3),
4-log scale removal of E. coli was achieved at an applied voltage of 0.2 V (against
SHE) and Fe2+ concentration of 0.3 mM. The production of •OH caused microbial
cell membrane destruction, which was identified as a prime reason for the inactiva-
tion of E. coli (Zhou et al. 2018).
In another investigation, when Ni-Pd was used as cathode catalyst in a dual
chamber MEC, the highest H2O2 yield of 233 mg L1 day1 was observed at an
imposed potential of 0.8 V (against SHE). When treated anolyte with the most
probable number of 5.2 106 (9.3 104) per 100 mL was circulated into the
cathodic chamber, 5 ( 1) log-scale bacterial removal was achieved in 3 days of
contact time (Gupta et al. 2020). These investigations showed that the secondary and
tertiary treatment of wastewater can be achieved using BES by removing organic
matter in the anodic chamber and pathogens in the cathodic chamber. Presently
considerable lab-scaled investigations have been conducted on in-situ H2O2 synthe-
sis in the cathodic chamber of BES. However, it was reported that for upscaling, the
H2O2 production rates were significantly lower than lab-scaled investigations. It
might be due to the overpotential losses as well as H2O2 decomposition, which
should be addressed in upcoming research. Also, reactor design aspects ranging
from cathode size, electrode spacing, catholyte, and anolyte volume need to be
optimized for H2O2 generation and a complete downstream treatment scheme for
anaerobically treated effluent needs to be taken up (Sim et al. 2018; Chung et al.
2020).
310 M. M. Ghangrekar et al.
Fig. 11.3 Bio-electrochemical system-based biosensor for volatile fatty acids monitoring
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 311
AD is one of the most proven and proficient techniques of energy recovery from the
waste; whereas, BES is a relatively newer concept and offers an advantage of
one-step bioelectricity recovery when used as MFC. The process of AD is suscep-
tible to various factors such as pH, VFA, and ammonia concentration. The BES can
be effectively used for the recovery of nutrients and removal of ECs or it can be
integrated with AD as biosensor for better process control. Unlike AD, for BES a
single large reactor is not suitable due to deteriorated electrical performance. Instead,
stacking of multiple smaller BES have been recommended and used by the
researchers for the wastewater treatment at field scale (Feng et al. 2014; Walter
et al. 2018).
The cost is yet another governing factor for the integration of AD and BES.
Typically in BES, more than 60% of the overall cost is associated with the electrodes
and separator (Rozendal et al. 2008). Metal electrodes have higher conductivity;
however, they are susceptible to corrosion in anaerobic conditions. Additionally,
metal electrodes lack high surface area and can be potentially toxic to few microor-
ganisms (Noori et al. 2020). For the pilot-scale and field-scale applications, low-cost
electrodes and separators need to be developed and integrated into the hybrid
system. At the same time, these components should be effective and durable for
long term operations. For the use as a separator, ceramic-based membranes show
better mechanical and structural properties than polymeric membranes at much
reduced cost; hence it can be an ideal choice for separators to be used in scaled up
BES (Ghadge and Ghangrekar 2015; Ieropoulos et al. 2015).
The application of BES also needs to be explored in the field-scale setups,
augmented with already operated anaerobic digesters to assess the potential of
312 M. M. Ghangrekar et al.
BES in real life scenarios in the upcoming research to help in confirming the
feasibility of the integration. Furthermore, for the integrated systems, the environ-
mental impact in terms of energy balance as well as material balance also needs to be
quantified via the use of life cycle analysis tools that will aid in considering the
environmental viability of the integrated system. As discussed earlier, BES have an
immense potential to complement AD for improving wastewater treatment effi-
ciency. However, more experience is to be gained by operating real life systems to
evaluate the efficacy of this integration and to come up with suitable medications in
system configurations to gain real advantage of this integration.
References
Inglesby AE, Fisher AC (2012) Enhanced methane yields from anaerobic digestion of Arthrospira
maxima biomass in an advanced flow-through reactor with an integrated recirculation loop
microbial fuel cell. Energ Environ Sci 5:7996–8006. https://doi.org/10.1039/c2ee21659k
Jadhav DA, Ghadge AN, Ghangrekar MM (2014) Simultaneous organic matter removal and
disinfection of wastewater with enhanced power generation in microbial fuel cell. Bioresour
Technol 163:328–334. https://doi.org/10.1016/j.biortech.2014.04.055
Jafary T, Daud WRW, Ghasemi M, Kim BH, Md Jahim J, Ismail M, Lim SS (2015) Biocathode in
microbial electrolysis cell; present status and future prospects. Renew Sustain Energy Rev 47:
23–33. https://doi.org/10.1016/j.rser.2015.03.003
Jain A, He Z (2018) “NEW” resource recovery from wastewater using bioelectrochemical systems:
moving forward with functions. Front Environ Sci Eng 12:1. https://doi.org/10.1007/s11783-
018-1052-9
Jeremiasse AW, Hamelers HVM, Buisman CJN (2010) Microbial electrolysis cell with a microbial
biocathode. Bioelectrochemistry 78(1):39–43. https://doi.org/10.1016/j.bioelechem.2009.
05.005
Jermakka J, Thompson Brewster E, Ledezma P, Freguia S (2018) Electro-concentration for
chemical-free nitrogen capture as solid ammonium bicarbonate. Sep Purif Technol 203:48–
55. https://doi.org/10.1016/j.seppur.2018.04.023
Jiang X, Shen J, Han Y, Lou S, Han W, Sun X, Li J, Mu Y, Wang L (2016) Efficient nitro reduction
and dechlorination of 2,4-dinitrochlorobenzene through the integration of bioelectrochemical
system into upflow anaerobic sludge blanket: a comprehensive study. Water Res 88:257–265.
https://doi.org/10.1016/j.watres.2015.10.023
Jin X, Angelidaki I, Zhang Y (2016) Microbial electrochemical monitoring of volatile fatty acids
during anaerobic digestion. Environ Sci Technol 50(8):4422–4429. https://doi.org/10.1021/acs.
est.5b05267
Jin X, Zhang Y, Li X, Zhao N, Angelidaki I (2017) Microbial electrolytic capture, separation and
regeneration of CO2 for biogas upgrading. Environ Sci Technol 51(16):9371–9378. https://doi.
org/10.1021/acs.est.7b01574
Jingura RM, Matengaifa R (2009) Optimization of biogas production by anaerobic digestion for
sustainable energy development in Zimbabwe. Renew Sustain Energy Rev 13(5):1116–1120.
https://doi.org/10.1016/j.rser.2007.06.015
Kahoush M, Behary N, Cayla A, Nierstrasz V (2018) Bio-Fenton and Bio-electro-Fenton as
sustainable methods for degrading organic pollutants in wastewater. Process Biochem 64:
237–247. https://doi.org/10.1016/j.procbio.2017.10.003
Kang HJ, Lee SH, Lim TG, Park JH, Kim B, Buffière P, Park HD (2021) Recent advances in
methanogenesis through direct interspecies electron transfer via conductive materials: a molec-
ular microbiological perspective. Bioresour Technol 322:124587. https://doi.org/10.1016/j.
biortech.2020.124587
Kaur A, Kim JR, Michie I, Dinsdale RM, Guwy AJ, Premier GC (2013) Microbial fuel cell type
biosensor for specific volatile fatty acids using acclimated bacterial communities. Biosens
Bioelectron 47:50–55. https://doi.org/10.1016/j.bios.2013.02.033
Kim BH, Park HS, Kim HJ, Kim GT, Chang IS, Lee J, Phung NT (2004) Enrichment of microbial
community generating electricity using a fuel-cell-type electrochemical cell. Appl Microbiol
Biotechnol 63:672–681. https://doi.org/10.1007/s00253-003-1412-6
Kim J, Song Y, Munussami G, Kim C, Jeon B (2015a) Recent applications of bioelectrochemical
system for useful resource recovery: retrieval of nutrient and metal from wastewater. Geosyst
Eng 18(4):173–180. https://doi.org/10.1080/12269328.2015.1037966
Kim T, An J, Jang JK, Chang IS (2015b) Coupling of anaerobic digester and microbial fuel cell for
COD removal and ammonia recovery. Bioresour Technol 195:217–222. https://doi.org/10.
1016/j.biortech.2015.06.009
Kiyasudeen S K, Ibrahim MH, Quaik S, Ahmed Ismail S (2016) An introduction to anaerobic
digestion of organic wastes. In: Prospects of organic waste management and the significance of
earthworms. Springer, Berlin, pp 1–21. https://doi.org/10.1007/978-3-319-24708-3
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 315
Kuntke P, Śmiech KM, Bruning H, Zeeman G, Saakes M, Sleutels THJA, Hamelers HVM,
Buisman CJN (2012) Ammonium recovery and energy production from urine by a microbial
fuel cell. Water Res 46(8):2627–2636. https://doi.org/10.1016/j.watres.2012.02.025
Kuntke P, Sleutels THJA, Rodríguez Arredondo M, Georg S, Barbosa SG, ter Heijne A, Hamelers
HVM, Buisman CJN (2018) (Bio)electrochemical ammonia recovery: progress and perspec-
tives. Appl Microbiol Biotechnol 102:3865–3878. https://doi.org/10.1007/s00253-018-8888-6
Lay C, Vo T, Lin P, Abdul P, Liu C, Lin C (2019) Anaerobic hydrogen and methane production
from low-strength beverage wastewater. Int J Hydrogen Energy 44(28):14351–14361. https://
doi.org/10.1016/j.ijhydene.2019.03.165
Lee DJ, Liu X, Weng HL (2014) Sulfate and organic carbon removal by microbial fuel cell with
sulfate-reducing bacteria and sulfide-oxidising bacteria anodic biofilm. Bioresour Technol 156:
14–19. https://doi.org/10.1016/j.biortech.2013.12.129
Lens PNL, Visser A, Janssen AJH, Hulshoff Pol LW, Lettinga G (1998) Biotechnological treatment
of sulfate-rich wastewaters. Crit Rev Environ Sci Technol 28(1):41–88. https://doi.org/10.1080/
10643389891254160
Li X, Angelidaki I, Zhang Y (2017) Salinity-gradient energy driven microbial electrosynthesis of
hydrogen peroxide. J Power Sources 341:357–365. https://doi.org/10.1016/j.jpowsour.2016.
12.030
Li S, Hua T, Yuan CS, Li B, Zhu X, Li F (2020) Degradation pathways, microbial community and
electricity properties analysis of antibiotic sulfamethoxazole by bio-electro-Fenton system.
Bioresour Technol 298:122501. https://doi.org/10.1016/j.biortech.2019.122501
Lin Y, Guo M, Shah N, Stuckey DC (2016) Economic and environmental evaluation of nitrogen
removal and recovery methods from wastewater. Bioresour Technol 215:227–238. https://doi.
org/10.1016/j.biortech.2016.03.064
Lin L, Xu F, Ge X, Li Y (2018) Improving the sustainability of organic waste management practices
in the food-energy-water nexus: a comparative review of anaerobic digestion and composting.
Renew Sustain Energy Rev 89:151–167. https://doi.org/10.1016/j.rser.2018.03.025
Liu H, Cheng S, Logan BE (2005) Production of electricity from acetate or butyrate using a single-
chamber microbial fuel cell. Environ Sci Technol 39(2):658–662. https://doi.org/10.1021/
es048927c
Liu L, Yuan Y, Li F-B, Feng C-H (2011a) In-situ Cr(VI) reduction with electrogenerated hydrogen
peroxide driven by iron-reducing bacteria. Bioresour Technol 102(3):2468–2473. https://doi.
org/10.1016/j.biortech.2010.11.013
Liu Y, Zhang Y, Quan X, Zhang J, Zhao H, Chen S (2011b) Effects of an electric field and zero
valent iron on anaerobic treatment of azo dye wastewater and microbial community structures.
Bioresour Technol 102(3):2578–2584. https://doi.org/10.1016/j.biortech.2010.11.109
Liu C, Sun D, Zhao Z, Dang Y, Holmes DE (2019) Methanothrix enhances biogas upgrading in
microbial electrolysis cell via direct electron transfer. Bioresour Technol 291:121877. https://
doi.org/10.1016/j.biortech.2019.121877
Logan BE, Hamelers B, Rozendal R, Schröder U, Keller J, Freguia S, Aelterman P, Verstraete W,
Rabaey K (2006) Microbial fuel cells: methodology and technology. Environ Sci Technol
40(17):5181–5192. https://doi.org/10.1021/es0605016
Ma B, Wang S, Cao S, Miao Y, Jia F, Du R, Peng Y (2016) Biological nitrogen removal from
sewage via anammox: recent advances. Bioresour Technol 200:981–990. https://doi.org/10.
1016/j.biortech.2015.10.074
Mao C, Feng Y, Wang X, Ren G (2015) Review on research achievements of biogas from anaerobic
digestion. Renew Sustain Energy Rev 45:540–555. https://doi.org/10.1016/j.rser.2015.02.032
Maurer M, Schwegler P, Larsen T (2003) Nutrients in urine: energetic aspects of removal and
recovery. Water Sci Technol 48(1):37–46. https://doi.org/10.2166/wst.2003.0011
McCartney DM, Oleszkiewicz JA (1991) Sulfide inhibition of anaerobic degradation of lactate and
acetate. Water Res 25(2):203–209. https://doi.org/10.1016/0043-1354(91)90030-T
316 M. M. Ghangrekar et al.
Miltner M, Makaruk A, Harasek M (2017) Review on available biogas upgrading technologies and
innovations towards advanced solutions. J Clean Prod 161:1329–1337. https://doi.org/10.1016/
j.jclepro.2017.06.045
Nancharaiah YV, Venkata Mohan S, Lens PNL (2016) Recent advances in nutrient removal and
recovery in biological and bioelectrochemical systems. Bioresour Technol 215:173–185. https://
doi.org/10.1016/j.biortech.2016.03.129
Noori M, Ghangrekar M, Mukherjee C, Min B (2019) Biofouling effects on the performance of
microbial fuel cells and recent advances in biotechnological and chemical strategies for miti-
gation. Biotechnol Adv 37(8):107420. https://doi.org/10.1016/j.biotechadv.2019.107420
Noori MT, Vu MT, Ali RB, Min B (2020) Recent advances in cathode materials and configurations
for upgrading methane in bioelectrochemical systems integrated with anaerobic digestion.
Chem Eng J 392:123689. https://doi.org/10.1016/j.cej.2019.123689
Peng Y, Zhu G (2006) Biological nitrogen removal with nitrification and denitrification via nitrite
pathway. Appl Microbiol Biotechnol 73:15–26. https://doi.org/10.1007/s00253-006-0534-z
Potter MC (1911) Electrical effects accompanying the decomposition of organic compounds. Proc
R Soc Lond Ser B Contain Pap Biol Character 84(571):260–276. https://doi.org/10.1098/rspb.
1911.0073
Qin M, Molitor H, Brazil B, Novak JT, He Z (2016) Recovery of nitrogen and water from landfill
leachate by a microbial electrolysis cell-forward osmosis system. Bioresour Technol 200:485–
492. https://doi.org/10.1016/j.biortech.2015.10.066
Rabaey K, Verstraete W (2005) Microbial fuel cells: novel biotechnology for energy generation.
Trends Biotechnol 23(6):291–298. https://doi.org/10.1016/j.tibtech.2005.04.008
Rabaey K, Van De Sompel K, Maignien L, Boon N, Aelterman P, Clauwaert P, De
Schamphelaire L, Pham HT, Vermeulen J, Verhaege M, Lens P, Verstraete W (2006) Microbial
fuel cells for sulfide removal. Environ Sci Technol 40(17):5218–5224. https://doi.org/10.1021/
es060382u
Rafiee A, Khalilpour KR, Prest J, Skryabin I (2021) Biogas as an energy vector. Biomass Bioenergy
144:105935. https://doi.org/10.1016/j.biombioe.2020.105935
Rozendal RA, Hamelers HVM, Rabaey K, Keller J, Buisman CJN (2008) Towards practical
implementation of bioelectrochemical wastewater treatment. Trends Biotechnol
26(8):450–459. https://doi.org/10.1016/j.tibtech.2008.04.008
Schievano A, Colombo A, Cossettini A, Goglio A, D’Ardes V, Trasatti S, Cristiani P (2018) Single-
chamber microbial fuel cells as on-line shock-sensors for volatile fatty acids in anaerobic
digesters. Waste Manag 71:785–791. https://doi.org/10.1016/j.wasman.2017.06.012
Semenec L, Franks A (2015) Delving through electrogenic biofilms: from anodes to cathodes to
microbes. AIMS Bioeng 2(3):222–248. https://doi.org/10.3934/bioeng.2015.3.222
Sevda S, Garlapati K, Naha S, Sharma M, Ray SG, Sreekrishnan TR, Goswami P (2020) Biosensing
capabilities of bioelectrochemical systems towards sustainable water streams: technological
implications and future prospects. J Biosci Bioeng 129(6):647–656. https://doi.org/10.1016/j.
jbiosc.2020.01.003
Sharma M, Kumar M (2020) Sulphate contamination in groundwater and its remediation: an
overview. Environ Monit Assess 192:74. https://doi.org/10.1007/s10661-019-8051-6
Shen J, Xu X, Jiang X, Hua C, Zhang L, Sun X, Li J, Mu Y, Wang L (2014) Coupling of a
bioelectrochemical system for p-nitrophenol removal in an upflow anaerobic sludge blanket
reactor. Water Res 67:11–18. https://doi.org/10.1016/j.watres.2014.09.003
Sheng K, Chen X, Pan J, Kloss R, Wei Y, Ying Y (2013) Effect of ammonia and nitrate on biogas
production from food waste via anaerobic digestion. Biosyst Eng 116(2):205–212. https://doi.
org/10.1016/j.biosystemseng.2013.08.005
Sim J, An J, Elbeshbishy E, Ryu H, Lee HS (2015) Characterization and optimization of cathodic
conditions for H2O2 synthesis in microbial electrochemical cells. Bioresour Technol 195:31–36.
https://doi.org/10.1016/j.biortech.2015.06.076
11 Integration of Bio-electrochemical Systems with Anaerobic Digestion 317
Zhang J, Zhang Y, Quan X, Li Y, Chen S, Zhao H, Wang D (2012b) An anaerobic reactor packed
with a pair of Fe-graphite plate electrodes for bioaugmentation of azo dye wastewater treatment.
Biochem Eng J 63:31–37. https://doi.org/10.1016/j.bej.2012.01.008
Zhang C, Su H, Baeyens J, Tan T (2014a) Reviewing the anaerobic digestion of food waste for
biogas production. Renew Sustain Energy Rev 38:383–392. https://doi.org/10.1016/j.rser.2014.
05.038
Zhang F, Li J, He Z (2014b) A new method for nutrients removal and recovery from wastewater
using a bioelectrochemical system. Bioresour Technol 166:630–634. https://doi.org/10.1016/j.
biortech.2014.05.105
Zhang S, Bao R, Lu J, Sang W (2018) Simultaneous sulfide removal, nitrification, denitrification
and electricity generation in three-chamber microbial fuel cells. Sep Purif Technol 195:314–
321. https://doi.org/10.1016/j.seppur.2017.12.027
Zhou S, Huang S, Li X, Angelidaki I, Zhang Y (2018) Microbial electrolytic disinfection process
for highly efficient Escherichia coli inactivation. Chem Eng J 342:220–227. https://doi.org/10.
1016/j.cej.2018.02.090
Zhu X, Ni J (2009) Simultaneous processes of electricity generation and p-nitrophenol degradation
in a microbial fuel cell. Electrochem Commun 11(2):274–277. https://doi.org/10.1016/j.elecom.
2008.11.023
Chapter 12
Use of Biogas for Electricity-Driven
Appliances
Afshin Davarpanah
Abstract Biogas can be considered as one of the primary renewable energy sources
to generate electricity regarding the grid connection and feed-in tariffs in industrial
plants. The present study is focused on introducing biogas systems on electricity
production and what environmental policies should be assessed to minimize the air’s
biogas component emission. Micro-gas turbine (MGT) systems, combined heat and
power (CHP) systems, solid oxide fuel cells (SOFC), and organic Rankine cycle
(ORC) systems are the most applicable renewable energy systems that can utilize
biogas. Combinations of technologies as a hybrid and novel system give the
engineers chance to optimize the biogas conversion to electricity for specified
industrial purposes. Moreover, environmental features of biogas emissions from
industrial plants were discussed and explained.
12.1 Introduction
Due to approaching depletion of the fossil fuels and out dependence on them in
recent decades, it is recommended to move towards renewable energy and green
technologies. Electricity generation is essential for industrial purposes and genera-
tion of electricity from green energy sources such as solar and wind energy should be
considered (Andalib-Bin-Karim et al. 2017; Hu et al. 2020; Ardebili 2020). Gener-
ating electricity from green energy sources is economically and environmentally
beneficial. Bioenergy resources are considered one of the preferable renewable
power generation resources due to their sustainable and productive clean electric
power (Chang and Zhao 2012; Effah and Boampong 2015; Qadrdan et al. 2018).
Right now, the electric power generation from renewable sources is about 8% and
A. Davarpanah (*)
Department of Mathematics, Aberystwyth University, Aberystwyth, UK
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 319
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_12
320 A. Davarpanah
Fig. 12.1 Solid oxide fuel cells for the conversion of biogas to electricity (Baldinelli et al. 2017)
Removal
Digestion Plant vessels
Feedstock COMP COMP
Air
Fuel
COMB
Removal
G COMP GT vessels
MC
Chiller
Heat NG
Demand
Exhuast
EVAP ECON Gas
HX
HRSG
Heat to Digester
Fig. 12.2 Micro-gas turbine (MGT) system for converting biogas to electricity (Kang et al. 2014)
electricity production. Therefore, the electricity and heat price was 0.1436 $/kWh
and 0.0738 $/kWh (Kang et al. 2014). Moreover, according to Kim et al. (2017),
electricity and heat production is estimated at 7.4 MWh/year and 8.5 MWh/year for
MGT systems in combined heat and power (CHP) and organic rankine cycle (ORC)
322 A. Davarpanah
Air
1073 K
Air
Fig. 12.3 Solid oxide fuel cells coupled with micro gas turbine and combined heat and power
system for the conversion of biogas to electricity (Wongchanapai et al. 2013)
systems. Combining the MGT and ORC systems has provided better results than the
MGT-CHP system (Kim et al. 2017).
Wongchanapai et al. (2013) developed a novel hybrid system containing solid
oxide fuel cells, micro gas turbine, and a combined heat and power system to convert
biogas to electricity (Fig. 12.3). In this hybrid system, an internal reformation unit
was implemented to convert the methane to hydrogen.
combined to increase the power generation efficiency. Thereby, waste gases have
been removed through the system circulation, and there are less hazardous emis-
sions. According to the measurements of biogas emissions, CO2 emissions range
from 0.1 to 0.4 kg CO2/kWh. Another biogas emission gas is NOx, which depends
on the nitrogen content of the produced biogas. One of the crucial factors in
environmental conservation is the leakage of hazardous materials that increase
global warming. As AD of waste has fugitive ammonia and greenhouse emissions,
efficient controlling methods to avoid leakages need to be designed (Sommer 1997;
Börjesson and Berglund 2006; Karapidakis et al. 2010; Senbayram et al. 2014;
Carreras-Sospedra et al. 2016).
References
Effah B, Boampong E (2015) Biomass energy: a sustainable source of energy for development in
Ghana. Asian Bull Energy Econ Technol 2(1):6–12
Hu X, Xie J, Cai W, Wang R, Davarpanah A (2020) Thermodynamic effects of cycling carbon
dioxide injectivity in shale reservoirs. J Petrol Sci Eng 195:107717. https://doi.org/10.1016/j.
petrol.2020.107717
Inayat A, Raza M, Ghenai C, Shanableh A, Said Z, Samman S, Al-Mansori A, Lazkani A (2019)
Simulation of anaerobic co-digestion process for the biogas production using ASPEN PLUS. In:
2019 Advances in science and engineering technology international conferences (ASET). IEEE,
pp 1–5. https://doi.org/10.1109/ICASET.2019.8714403
Kang JY, Kim TS, Hur KB (2014) Economic evaluation of biogas and natural gas co-firing in gas
turbine combined heat and power systems. Appl Therm Eng 70(1):723–731. https://doi.org/10.
1016/j.applthermaleng.2014.05.085
Karapidakis ES, Tsave AA, Soupios PM, Katsigiannis YA (2010) Energy efficiency and environ-
mental impact of biogas utilization in landfills. Int J Environ Sci Technol 7(3):599–608. https://
doi.org/10.1007/BF03326169
Kim S, Sung T, Kim KC (2017) Thermodynamic performance analysis of a biogas-fuelled micro-
gas turbine with a bottoming organic rankine cycle for sewage sludge and food waste treatment
plants. Energies 10(3):275. https://doi.org/10.3390/en10030275
Kiran EU, Stamatelatou K, Antonopoulou G, Lyberatos G (2016) Production of biogas via
anaerobic digestion. In: Handbook of biofuels production: processes and technologies, 2nd
edn. Woodhead, pp 259–301. https://doi.org/10.1016/B978-0-08-100455-5.00010-2
Liu X, Gao X, Wang W, Zheng L, Zhou Y, Sun Y (2012) Pilot-scale anaerobic co-digestion of
municipal biomass waste: focusing on biogas production and GHG reduction. Renew Energy
44:463–468. https://doi.org/10.1016/j.renene.2012.01.092
Makareviciene V, Sendzikiene E, Pukalskas S, Rimkus A, Vegneris R (2013) Performance and
emission characteristics of biogas used in diesel engine operation. Energy Convers Manag 75:
224–233. https://doi.org/10.1016/j.enconman.2013.06.012
Mao C, Feng Y, Wang X, Ren G (2015) Review on research achievements of biogas from anaerobic
digestion. Renew Sustain Energy Rev 45:540–555. https://doi.org/10.1016/j.rser.2015.02.032
Pöschl M, Ward S, Owende P (2010) Evaluation of energy efficiency of various biogas production
and utilization pathways. Appl Energy 87(11):3305–3321. https://doi.org/10.1016/j.apenergy.
2010.05.011
Qadrdan M, Jenkins N, Wu J (2018) Smart grid and energy storage. In: McEvoy’s handbook of
photovoltaics. Academic, pp 915–928. https://doi.org/10.1016/B978-0-12-809921-6.00025-2
REN21 (2018) Renewables 2018 global status report. REN21 Secretariat, Paris. https://www.ren21.
net/wp-content/uploads/2019/05/GSR2018_Full-Report_English.pdf
Saadabadi SA, Thattai AT, Fan L, Lindeboom RE, Spanjers H, Aravind PV (2019) Solid oxide fuel
cells fuelled with biogas: potential and constraints. Renew Energy 134:194–214. https://doi.org/
10.1016/j.renene.2018.11.028
Senbayram M, Chen R, Wienforth B, Herrmann A, Kage H, Mühling KH, Dittert K (2014)
Emission of N2O from biogas crop production systems in Northern Germany. BioEnergy Res
7(4):1223–1236. https://doi.org/10.1007/s12155-014-9456-2
Sommer SG (1997) Ammonia volatilization from farm tanks containing anaerobically digested
animal slurry. Atmos Environ 31(6):863–868. https://doi.org/10.1016/S1352-2310(96)00250-6
Wongchanapai S, Iwai H, Saito M, Yoshida H (2013) Performance evaluation of a direct-biogas
solid oxide fuel cell-micro gas turbine (SOFC-MGT) hybrid combined heat and power (CHP)
system. J Power Sources 223:9–17. https://doi.org/10.1016/j.jpowsour.2012.09.037
Wu B, Zhang X, Shang D, Bao D, Zhang S, Zheng T (2016) Energetic-environmental-economic
assessment of the biogas system with three utilization pathways: combined heat and power,
biomethane and fuel cell. Bioresour Technol 214:722–728. https://doi.org/10.1016/j.biortech.
2016.05.026
Chapter 13
Syngas Fermentation for Bioenergy
Production: Advances in Bioreactor
Systems
A. Sinharoy (*)
National University of Ireland Galway, Galway, Ireland
Department of Biosciences and Bioengineering, Indian Institute of Technology Guwahati,
Guwahati, Assam, India
e-mail: [email protected]
K. Pakshirajan
Department of Biosciences and Bioengineering, Indian Institute of Technology Guwahati,
Guwahati, Assam, India
P. N. L. Lens
National University of Ireland Galway, Galway, Ireland
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 325
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_13
326 A. Sinharoy et al.
13.1 Introduction
The majority of today’s energy demand is fulfilled by energy supplied from fossil
fuel resources, as the population and industrial growth progresses, the demand for
energy is going to increase furthermore (Johnsson et al. 2019; Sinharoy et al. 2020a).
This is one of the main reason for the exhaustion of fossil fuel reserve, which have
created many global problems and can potentially leads to future energy crisis (Roy
et al. 2016). Apart from that, environmental problems associated with fossil fuel
burning and their effect on human health have created worldwide concern and have
generated demand for clean energy (Chen et al. 2017). These factors combined with
the awareness about global warming have shifted our focus towards safeguarding the
environment and creating a more sustainable society by fulfilling our energy
demands from renewable sources.
The use of renewable energy sources can provide an alternative to the fossil fuel
and simultaneously avoid the environmental degradation associated with fossil fuel
to a larger extent (Soeiro and Dias 2020; Jurasz et al. 2020; Sinsel et al. 2020).
Biofuels, such a renewable energy source, have received a lot of attention during the
recent years. Biofuels are available mainly in liquid form, such as bioethanol,
biobutanol and biodiesel, or in gaseous form, such as biogas (mainly methane) and
biohydrogen (Bórawski et al. 2019). These fuels are mainly derived from biomass as
the raw material or produced via biological or biochemical conversion processes
(Sinharoy et al. 2020a). Due to the conflict with food crops, limited availability of
agricultural land and large water requirements for production of first generation
biofuels, the recent focus is towards the production of biofuels utilizing waste
resources (Bhatia et al. 2017). Solid as well as liquid wastes such as municipal
solid waste, lignocellulosic biomass, agricultural and forest residues and different
industrial wastewaters have been used for production of various types of biofuels
(Hassan et al. 2019). In contrast, the potential of gaseous industrial wastes for
production of biofuel or other useful products has not yet been fully explored. In
this context, the use of synthesis gas or syngas as a low cost substrate for the
production of biofuel or other value added products is attractive. This gas fermen-
tation approach is another way of producing second generation biofuels from waste
biomass.
Synthesis gas is a mixture of carbon monoxide (CO), hydrogen (H2), carbon
dioxide (CO2), nitrogen (N2) and some higher hydrocarbons. The CO concentration
in syngas ranges from 5 to 60%. CO can be steam reformed to enrich the H2 content
of the syngas (Couto et al. 2013). Biomass gasification, which is the main route for
syngas production, is an endothermic process and requires heat energy as an input as
is carried out at 750–800 C (Molino et al. 2016). Lignocellulosic biomass and even
municipal solid waste can be feedstocks for biomass gasification (Ahmad et al.
2016). For an efficient gasifier operation, a certain amount of feedstock homogeneity
should be maintained. If the feedstock is heterogeneous then the product composi-
tion may vary widely (Sikarwar et al. 2016). To overcome this bottleneck,
pre-treatment and post-treatment steps are included in the gasification process,
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 327
which increase the operation cost of the process (Cheah et al. 2016). Homogeneous
lignocellulosic materials, such as timber industry waste, paper mill waste and
agricultural residues can be used as feedstock for biomass gasification (Sikarwar
et al. 2016; Burra and Gupta 2018). Although, municipal solid waste can be a
suitable substrate for biomass gasification as well, it is still under consideration as
it also contains a variety of other (non-combustible) materials (Zheng et al. 2018).
Syngas can be used as the substrate for production of liquid fuel either by the
thermochemical catalytic route with a metallic catalyst or by the biochemical route
with acetogenic microorganisms (Griffin and Schultz 2012). A number of thermo-
chemical catalytic processes such as Fischer–Tropsch (FT) synthesis and the water
gas shift reaction have been applied for converting syngas to valuable products.
These operate at high temperature and pressure, and require costly metal catalysts
(Subramani and Gangwal 2008). Though the thermo-catalytic process has been
commercialized and works successfully in installed plants, it has several disadvan-
tages such as high operating costs, limited choice of metallic catalysts, perishable
nature of the catalyst, huge requirement of heat energy and environmental concerns
(Griffin and Schultz 2012). Furthermore, these catalyst based processes are often
prone to catalyst poisoning due to the emanation of unwanted gases, mainly hydro-
gen sulfide and other inhibitory substances (Subramani and Gangwal 2008). These
disadvantages have resulted in the need for alternative technologies, including
biochemical conversion of syngas. Biological methods offer certain advantages
such as high tolerance to trace contaminants, high product specificity, in addition
to being sustainable, environmental friendly and cost effective (Henstra et al. 2007).
Some acetogens are capable of producing alcohols along with volatile fatty acids
from syngas (Table 13.1). These bacteria are obligate anaerobes and can utilize
CO/CO2 as the carbon and energy source (Henstra et al. 2007; Mohammadi et al.
2011). Examples of these solventogenic bacteria are Clostridium ljungdahlii, Clos-
tridium carboxidivorans, Clostridium autoethanogenum, Clostridium ragsdalei,
Alkalibaculum bacchi, Eubacterium limosum, Butyribacterium methylotrophicum,
Clostridium thermoaceticum and Clostridium formicoaceticum (Mohammadi et al.
2011; Bengelsdorf et al. 2013). Most of these microorganisms can grow on multiple
carbon sources, but their potential to utilize gaseous substrates such as CO, CO2 and
H2 and to produce useful products is commercially attractive. Some strains such as
Clostridium ragsdalei and Clostridium propionicum produce propanol from syngas
(Isom et al. 2015). The biochemical pathway used by these bacteria for the produc-
tion of ethanol is the reductive acetyl-CoA pathway or Wood-Ljungdahl pathway
(Mohammadi et al. 2011). This pathway is present in several organisms including
acetogenic bacteria (Schuchmann and Müller 2016) and methanogenic archaea
Table 13.1 Alcohol production from syngas in different bioreactor systems
328
Clostridium CSTR 37 6.1 150 20% CO, 5% H2, 15% NR 9.6 g/L Acetic acid Maddipati
ragsdalei CO2, 60% N2 et al. (2011)
Clostridium TBR 37 5.8 NR 38% CO, 28.5% CO2, NR 5.7 g/L Acetic acid Devarapalli
ragsdalei 28.5% H2, 5% N2 et al. (2016)
Clostridium CSTR 37 6.0 150 20% CO, 5% H2, 15% NR 25.26 g/L 2-propanol, Kundiyana
ragsdalei (pilot scale CO2, 60% N2 Acetic acid, et al. (2010)
100L) 1-butanol
Alkalibaculum BSB—fed 37 8.0–8.5 150 20% CO, 15% CO2, NR 1.75 g/L Acetic acid Liu et al.
bacchi batch mode 5% H2, 60% N2 and (2012)
40% CO, 30% CO2,
30% H2
Alkalibaculum CSTR 37 8.0 NR 20% CO, 5% H2, 15% Corn steep 8 g/L Acetic acid, Liu et al.
bacchi CO2, 60% N2 liquor n-propanol (2014)
and n-butanol
Anaerobic mixed BSB 37 6.0 100 PCO ¼ 2 kg/cm2 NR 2.2 g/L Acetic acid Singla et al.
consortium (2014)
NR, not reported; NA, not applicable; BSB, batch serum bottle; CSTR, continuous stirred tank reactor; HRPBR, horizontal rotating packed bed reactor; BCR,
bubble column rector; TBR, trickle bed reactor
Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems
329
330 A. Sinharoy et al.
Fig. 13.1 Wood-Ljungdahl pathway or reductive acetyl-CoA pathway utilized for syngas fermen-
tation. Abbreviations: ACS, acetyl-CoA synthetase; AckA acetate kinase; [CO], enzyme-bound
carbon monoxide; CODH, carbon monoxide dehydrogenase; CoFeSP, corrinoid sulfur protein;
FDH, formate dehydrogenase; FHS, formyl-THF synthetase; [H], reducing equivalents; HS-CoA,
coenzyme A; MET, methyltransferase; Pi, inorganic phosphate; PTA, phosphotransacetylase; THF,
tetrahydrofolic acid [reused from Kiefer et al. 2021; Copyright (2021) with permission from
Elsevier]
(Borrel et al. 2016). It consists of two branches, eastern branch and western branch
(Fig. 13.1). The eastern branch has a number of reductive steps where CO2 is
reduced to produce the methyl group of acetyl-CoA. Whereas in the western branch,
CO obtained from CO2 or directly from the source forms the carbonyl group for the
acetyl-CoA synthesis (Diender et al. 2015). There are two ways to generate reducing
equivalents by this pathway. Most commonly, a hydrogenase enzyme supplies the
reducing equivalents from H2 as per the following reaction (Eq. 13.1) (Diender et al.
2015):
oxidation of CO to CO2 as per the following reaction (Eq. 13.2) (Diender et al.
2015):
Any other
Temp Agitation carbon H2 production
Microorganism Reactor C pH (rpm) Feed gas composition substrate rate /yield Reference
Rhodospirillum CSTR 30 6.5 NR PCO ¼ 0.55 atm Acetate 16 mmol H2/h Najafpour et al.
rubrum (2003)
Rhodospirillum CSTR 30 6.9 350–800 55% CO, 15% H2, 20% Malic acid/ 9.6 mmol H2/h Ismail et al.
rubrum Ar, 10% CO2 sodium (2008)
acetate
Rhodospirillum CSTR 30 6.9 150–500 55% CO, 15% H2, 20% Malic acid/ 80% mol/mol Younesi et al.
rubrum Ar, 10% CO2 sodium of CO (2008)
acetate
Rhodospirillum CSTR 30 7.0 400–750 17.2% CO, 16.3% Acetate 0.75 mol H2/ Do et al. (2007)
rubrum CO2, 8.8% H2, 56.0% day
N2
Thermoanaerobacter BSB 60 7.5 NR 100% CO NR 45 mmol/L Weghoff and
kivui Müller (2016)
Rubrivivax BSB 30 7.0 250 20% CO Sodium 0.9 mmol/ min/ Vanzin et al.
gelatinosus malate g cell (2002)
Rhodopseudomonas BSB 30 7.0 NR 60% CO, 10% CO2, Sodium 7 mmol/L Pakpour et al.
palustris P4 20% H2, 10% Ar acetate (2014)
Rhodopseudomonas CSTR 30 7.0 500–700 20% CO, 80% Ar NR 41 mmol/g Oh et al. (2005)
palustris P4 cell/h
Citrobacter sp. Y19 Two stage Batch 30 7 300–700 20% CO, 80% Ar Sucrose 20 mmol/g Jung et al.
stirred tank reactor cell/h (2002)
Carboxydothermus GLR 70 6.8–7.2 NR 100% CO Bacto- 95% mol/mol Haddad et al.
hydrogenoformans peptone of CO (2014)
Carboxydothermus HFMBR 60–70 6.8–7.0 Nil 100% CO NR 92% mol/mol Zhao et al.
hydrogenoformans of CO (2013)
A. Sinharoy et al.
13
Proper design and operation of a bioreactor are important considerations for suc-
cessful scale up of syngas bioconversion to useful products (Yasin et al. 2015). The
key parameters necessary while selecting a suitable bioreactor are mainly related to
gas-liquid mass transfer, which is governed by agitation speed, impeller design,
power consumption, temperature, pressure conditions and bioreaction kinetics
(Munasinghe and Khanal 2010a; Yasin et al. 2015). Both batch and continuous
Table 13.3 Methane production from syngas in different bioreactor systems
336
Any
other
Temp Agitation carbon CH4 production
Inoculum Reactor C pH (rpm) Feed gas composition substrate rate /yield Reference
Anaerobic granular sludge GLR 35.2 7.1 NA CO loading ¼ 15–122- NR 0.49–4.77 mmol/ Guiot et al.
mmol/g VSS/day g VSS/day (2011)
Enriched TBR 37, 60 7.0 NA 45% H2, 25% CO2, NR 30–70% CH4 in Asimakopoulos
mixed microbial consortia 20% CO, 10% N2 biogas et al. (2021a)
Enriched thermophilic TBR 60 7.0 NA 45% H2, 25% CO2, NR 17.6 0.6 mmol/ Asimakopoulos
mixed microbial consortium 20% CO, 10% N2 L/h et al. (2021b)
Anaerobic sludge CSTR 37, 55 8.5 400–500 10% CO2, 40% CO, 50% NR 66–73% CH4 in Li et al. (2020)
H2 biogas
Anaerobic granular sludge UASB 37 7.28–7.65 NA 100%, 2.5 L CO/day Glucose 1709–2628 mL/ Jing et al.
L/day (2017)
Anaerobic sludge RMBR 55 7.0 NA 55% CO, 20% H2, 10% NR 186 mL/L/day Westman et al.
CO2 (2016)
Anaerobic sludge TBR 37, 60 7.0 NA 45% H2, 25% CO2, 20% NR 8.49 mmol/L/h Asimakopoulos
CO, 10% N2 et al. (2020)
Anaerobic sludge TBR 25, 50 7.0 NA 80% H2, 20% CO2 NR 13.1 m3/m3/day, Strübing et al.
97.5% CH4 in (2018)
biogas
Anaerobic sludge TBR 55 7.0 NA 80% H2, 20% CO2 NR 15.4 m3 CH4/m3/ Strübing et al.
day (2017)
Enrichment anaerobic Sealed 37, 60 7.2 100 PH2 ¼ 1 atm, NR 18.8 mmol/g Grimalt-
cultures flasks PCO ¼ 0.2–0.8 atm VSS/day Alemany et al.
(2020)
Anaerobic granular sludge BSB 35 7.5 200 20–100% CO NR 1 mmol/g Sancho Navarro
VSS/day et al. (2016)
A. Sinharoy et al.
13
Thermophilic mixed culture FMBR 55 8.0 NA 20% H2, 55% CO, 10% NR 34.41 mmol/L/ Chandolias
CO2 day et al. (2019)
Sewage sludge CSTR- 37 7.6 NA 2% H2, 8% NR 1322 mL/day Wang et al.
HFM CO biogas produc- (2013)
tion, 98.8% CH4
Anaerobic mixed cultures BSB 70 7.5 NA 80% H2, 5% CO, 15% NR 78.6–89.5 mL/ Bu et al. (2018)
CO2 day
Co-culture of CSTR 65 7.2 NR 66.66% H2, 33.33% CO Acetate 4.91 mmol/L/h Diender et al.
Carboxydothermus (2018)
hydrogenoformans and
Methanothermobacter
thermoautotrophicus
NR, not reported, NA, not applicable; BSB, batch serum bottle; CSTR, continuous stirred tank reactor; RMBR, reverse membrane bioreactor; FMBR, floating
membrane bioreactor; GLR, gas lift reactor; UASB, upflow anaerobic sludge blanket reactor; TBR, trickle bed reactor; HFM, hollow fiber membranes
Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems
337
338 A. Sinharoy et al.
operation modes using a few bioreactor configurations have been applied for syngas
fermentation to useful products. Continuous stirred-tank reactor (CSTR) is the most
commonly used bioreactor system for syngas fermentation, but other types of
reactors such as bubble column reactor (BCR), trickling bed reactor (TBR), packed
bed bioreactor (PBR), gas lift reactor (GLR) and membrane bioreactor (MBR)
(primarily hollow fibre membrane bioreactor) have also been examined under
different modes of operation and experimental conditions (Asimakopoulos et al.
2018). Table 13.4 compares the commonly used reactors for gas fermentation along
with their volumetric gas to liquid mass transfer coefficient (KLa) values. The
different types of bioreactor configurations studied for syngas fermentation to useful
products are depicted in Fig. 13.2 and discussed further in the following
sub-sections. Table 13.5 overviews the relative advantages and drawbacks of the
different types of bioreactors for syngas fermentation.
The continuous stirred tank reactor (CSTR; Fig. 13.2a) is the most commonly used
bioreactor type employed for syngas fermentation research as it offers good mixing.
Mixing of gaseous substrates in the liquid media inside the reactor is achieved by
impellers and baffles which enhance mass transfer between the substrate and micro-
organisms inside the reactor (Munasinghe and Khanal 2010b). The hydrodynamic
shear generated by a high level of agitation or mixing creates small sized bubbles
from bigger ones, which increases the interfacial area for mass transfer and subse-
quent enhancement in the bioavailability of gaseous substrates (Bredwell et al.
1999). Besides, the slow rising velocity of the finer gas bubbles lead to a prolonged
gas retention time in the liquid medium, resulting in high mass transfer rates. In this
type of reactor, gaseous substrate is supplied continuously and liquid medium
containing the nutrients is fed into the bioreactor to support microbial growth and
metabolism (Bredwell et al. 1999; Sinharoy et al. 2020a). Products formed in the
reactor by microbial conversion are drawn out from the system at the same flow rate
as the feed to maintain steady state profiles under continuous mode of operation.
The effect of different operational parameters, such as agitation speed, medium
flow rates and composition as well as syngas flow rate on ethanol and acetate
production by Clostridium ljungdahlii was studied using a CSTR (Aghbashlo
et al. 2016). The sustainability and renewability of the process was also evaluated
by a thermodynamic model. From the calculated model parameters, the maximum
exergetic productivity index was 6.82 and 6.90 using, respectively, the conventional
exergy and eco-exergy concepts at an optimum condition of 450 rpm agitation
speed, 0.55 mL/min liquid medium flow rate and 8 mL/min syngas (55% CO,
20% H2, 10% CO2 and 15% Ar) volumetric flow rate. A six fold increase in ethanol
concentration by Clostridium ragsdalei was achieved in a pilot scale (100 L) stirred
tank reactor compared with that in a serum bottle (Kundiyana et al. 2011). The
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 339
Table 13.4 Commonly used bioreactor systems for biological CO conversion and their volumetric
mass transfer coefficients
Volumetric
Agitation mass transfer
speed coefficient KLa Gaseous
Reactor type (rpm) (h–1) feed Products Reference
Continuous 200 14.2 CO NR Bredwell
stirred tank et al. (1999)
reactor
(CSTR)
CSTR 300 35 Syngas NR Bredwell
et al. (1999)
CSTR 450 101 Syngas NR Bredwell
et al. (1999)
CSTR 500 86.4 CO H2 ¼ 9.6 mmol/h Ismail et al.
(2008)
CSTR 450 153 CO NR Ungerman
and Heindel
(2007)
CSTR 600 154.8 CO NR Riggs and
Heindel
(2006)
CSTR 700 192.8 Syngas NR Kapic et al.
(2006)
CSTR 300 14.9 CO NR Klasson
et al. (1993)
CSTR 400 21.5 CO NR Klasson
et al. (1993)
CSTR 700 35.5 CO NR Klasson
et al. (1993)
Trickle bed NA 137 Syngas NR Cowger
reactor et al. (1992)
(TBR)
TBR NA 121 Syngas NR Bredwell
et al. (1999)
Bubble col- NA 72 CO Acetate ¼ Chang et al.
umn reactor 15.6–90 mM, butyrate (2001)
(BCR) ¼ 0.7–1.3 mM
Hollow fiber NA 400 CO NR Munasinghe
membrane and Khanal
reactor (2012)
(HFMR)
HFMR NA 1096.2 CO Acetate ¼ 8.2 g/L, Shen et al.
ethanol ¼ 23.93 g/L, (2014a)
butanol ¼ 0.45 g/L
HFMR NA 385 CO NR Lee et al.
(2012)
Packed bed NA 21 Syngas NR Bredwell
reactor et al. (1999)
(continued)
340 A. Sinharoy et al.
ethanol concentration reached a maximum value of 25.26 g/L along with the
co-products 2-propanol, 1-butanol and acetic acid.
A two-stage CSTR for ethanol production from syngas was studied using Clos-
tridium carboxidivorans, in which the pH of the first reactor was maintained at pH
6 for acidogenesis and the second one was maintained at pH 5 for solventogenesis
(Abubackar et al. 2018). This two-stage approach achieved a high biomass concen-
tration in the first stage and increased the ethanol production in the second stage.
Bioconversion of CO rich syngas to hydrogen using a CSTR with the anaerobic
photosynthetic bacterium Rhodospirillum rubrum, capable of performing the bio-
logical water-gas shift reaction, was studied (Younesi et al. 2008). The best perfor-
mance in terms of hydrogen production rate and conversion efficiency of 16 ( 1.1)
mmol/g cell/h and 87 ( 2.4) %, respectively, was achieved at an agitation speed and
gas flow rate of, respectively, 500 rpm and 14 mL/min. In order to achieve a high gas
liquid mass transfer a CSTR equipped with a dual-impeller agitation and
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 341
Fig. 13.2 Schematic of various commonly used bioreactor systems for syngas fermentation. (a)
Stirred tank reactor, (b) bubble column reactor, (c) packed bed reactor, (d) gas lift reactor and (e)
membrane bioreactor
microsparger system was utilized (Najafpour et al. 2003). A very high H2 production
rate of 16 mmol/g cell/h and 80% conversion efficiency on CO by R. rubrum at a
syngas flow rate of 14 mL/min was achieved. The maximum cell density during
steady state operation reached 1.3–1.4 g/L with a specific growth rate of 0.0225/h.
A bubble column reactor (Fig. 13.2b) offers a high gas-liquid mass transfer due to
the high surface area of bubbles and increased turbulence as the bubbles rise through
the liquid column. Small-sized finer bubbles produced in this type of reactor have
high surface-to-volume ratios and low rise velocities through the liquid, resulting in
improved contact times in the reactor. The short gas residence time and high pressure
drop in the bubble reactor are some drawbacks limiting its scale-up and commer-
cialization. Other drawbacks of this bioreactor type include back-mixing and
342 A. Sinharoy et al.
Table 13.5 Relative advantages and disadvantages of different bioreactors used for syngas
fermentation
Reactor type Advantages Disadvantages
Continuous • Simple to operate • High agitation causes shear
stirred tank • Mixing improves gas-liquid mass transfer stress on microbes
reactor • Gaseous substrates are more accessible to • Large power requirement
microbes due to breaking of large bubbles • High operating cost
into finer ones by increasing rotational speed
of impellers
Bubble col- • High gas-liquid mass transfer rates • Back mixing and coalescence
umn reactor • Low operating and maintaining cost
• Low shear stress
Trickle bed/ • Gas and liquid flow could be either in • Clogging of bed
packed bed co-current or counter-current direction
reactor • No need of mechanical agitation
• Power requirement is less
Membrane • Gases easily diffuses through wall of the • Not yet commercially used for
bioreactor membranes CO conversion or syngas fer-
• Membrane wall provides excellent support mentation
for microbial growth and improves biomass • High installation cost
retention • Clogging problems
• High product yield, high reaction rate, high
tolerance to toxic elements
Gas lift • Mechanical simplicity • Not suitable for viscous broths
reactor • Low energy requirement • Minimum process volume
• No focal points for energy dissipation • Dead zones inside the reactor
• Low shear stress • Insufficient mixing at high
• Suitable for process with variable gas feed- biomass densities
ing requirements
Moving bed • Good biomass retention • Prolonged startup time
biofilm • High surface area for biofilm formation • Loss of biomass support over
reactor • Auto-control of biomass due to shear and time
tear during mixing • Troubleshooting is difficult
• Less area requirement due to complicated design
coalescence in the bubble column (Datar et al. 2004). But the main attraction of this
type of reactor is the ease of obtaining the desired amount of biomass in the reactor.
Syngas fermentation by Clostridium carboxidivorans was studied in a continu-
ously operated bubble column bioreactor (4.5 L volume): ethanol, butanol, and
acetic acid yields were 0.15, 0.075 and 0.025 g/mol CO, respectively, under steady
state condition (Rajagopalan et al. 2002). The authors further suggest different
approaches such as cell recycle to increase the cell concentration without altering
the ethanol yield, H2 addition in the gas stream to improve the CO conversion to
ethanol and optimization of nutrient supplements to the medium to further improve
the bioreactor performance.
Amos (2004) evaluated the performance of a bubble column reactor for CO
bioconversion to H2 with or without gas recycle. The reactor with added gas recycle
loop performed better with an overall conversion efficiency of over 50% for a 10:1
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 343
The trickle-bed bioreactor or trickling filter (Fig. 13.2c) contains a tubular reactor
with solid support for biomass to grow and attach to. The cells can either be
immobilized on the solid packing or suspended in the liquid medium (Bredwell
et al. 1999). Generally, such bioreactor system is operated under counter current flow
of liquid to gas, where gaseous substrate rises upwards and the water flows down-
wards through the packed bed (Stoll et al. 2020). Special care needs to be taken to
maintain a low water flow rate to prevent flooding in the column. The liquid flow is
mainly provided to keep the cells moist along with nutrient supplementation for cell
growth (Sinharoy et al. 2020c). Various packing materials, including wood, acti-
vated carbon, lava rock, plastic and porous ceramic supports have been tested for
biofilm formation (Amos 2004). The primary advantage of this bioreactor configu-
ration is the improved gas transfer area with a minimum pressure drop. Also, it is
easy to enhance the mass transfer by controlling the liquid flow rates (Sinharoy et al.
2020c).
The effect of the reactor packing material on syngas fermentation by the photo-
synthetic bacterium Rubrivivax gelatinosus was examined using two geometrically
similar reactors, but with different volumes (1 and 5 L) and non-porous glass beads
of two different diameters (3 and 6 mm) as the support material (Wolfrum and Watt
2001). The CO conversion efficiency with the glass beads of 3 mm diameter at
different superficial liquid velocities was better than with the 6 mm diameter glass
beads, which was attributed to a high mass transfer with the smaller diameter
particles owing to their high surface area to volume ratio. However, the reactor
volume did not significantly affect the CO conversion, indicating consistency in
performance of this bioreactor configuration which is helpful in predicting its
performance.
A trickle bed bioreactor with 6 mm size soda lime glass beads as packing material
was studied for ethanol fermentation from syngas (Devarapalli et al. 2016). CO
inhibition could be overcome by a high amount of biomass available in the form of
biofilm formed over the beads, resulting in 1.9 times enhanced H2 uptake and
conversion. The final ethanol and acetic acid concentrations inside the reactor
reached a maximum value of 5.7 and 12.3 g/L, respectively. It was also found that
the co-current mode of gas and liquid flow in the reactor reduced the gas bypass and
reactor flooding problems that were encountered in counter-current operation.
Apart from passive immobilization methods to grow biomass on support mate-
rials, active immobilization techniques to entrap the anaerobic biomass inside
polyvinyl alcohol (PVA) or sodium alginate beads have also been studied (Kumar
et al. 2018). PVA addition to the sodium alginate beads improves their strength and
344 A. Sinharoy et al.
recycle potential for long term use. High CO utilization (90%) efficiencies were
achieved using a CO fed PBR operated at an HRT of 48 h. The superior performance
of the PBR containing immobilized beads is attributed to the excellent gas-liquid
mass transfer due to counter current flow of the liquid and the feed gas. Moreover,
the use of biomass immobilization in PVA-alginate beads can prevent CO toxicity to
the anaerobic biomass by avoiding direct exposure of the biomass to high CO
concentrations (Kumar et al. 2018).
Methane production from syngas using a trickle bed reactor was studied to
understand the effect of temperature on the process efficiency (Asimakopoulos
et al. 2020). Thermophilic conditions (60 C) were best suited for methane produc-
tion with a maximum methane productivity of 8.49 mmol/L/h. The microbial
community change at mesophilic conditions resulted in high accumulation of acetate
in the reactor and lowering of the final methane productivity. The absence of
acetoclastic methanogens in both the mesophilic and thermophilic reactor indicates
that the methane production occurred through the hydrogenotrophic route. A scale
up study using a 7.5 L volume reactor was also carried out during which a very high
methane productivity of 17.6 mmol/L/h was achieved at a 0.33 h empty bed
residence time (Asimakopoulos et al. 2021b). The reactor showed a maximum of
100% H2 and 92.4% CO utilization without any process inhibition during the
continuous bioreactor operation.
A gas lift reactor (GLR; Fig. 13.2d) is similar to the bubble column reactor, but it
differs by the fact that it contains a draft tube, through which syngas flows (Sinharoy
et al. 2020b). The draft tube is either an inner tube (called gas lift bioreactor with an
internal loop) or an external tube (called gas lift bioreactor with an external loop)
which improves circulation and gas liquid mass transfer and equalizes shear forces in
the reactor (Negi et al. 2020; Sinharoy et al. 2020c). The major advantages of a gas
lift bioreactor are its simple design with no moving parts or agitators, low energy
requirement as well as homogeneous distribution of nutrients and shear force
(Riegler et al. 2019). The gas stream facilitates the exchange of material between
the gas phase and the liquid phase, thus enhancing effective mass transfer. For
example, a very high mass transfer coefficient of 91.08/h for CO was reported in a
gas lift reactor combined with a bubble diffuser compared to other reactor configu-
rations (Munasinghe and Khanal 2010b).
A CO and H2 fed GLR with a high gas-liquid mass transfer coefficient
(kLa ¼ 80.28/h) was used to enrich acetogenic microorganisms from different
animal faeces to isolate strains capable of syngas fermentation (Park et al. 2013).
Ten out of 42 strains isolated from this enriched animal faeces were able to utilize
CO/H2 and produce alcohols and volatile fatty acids. The enriched consortium
isolated from chicken faecal samples showed the highest product formation from
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 345
syngas, i.e. 711 mg/L acetate, 498 mg/L ethanol, 389 mg/L butyrate and 251 mg/L
butanol (Park et al. 2013).
CO conversion to H2 by Carboxydothermus hydrogenoformans in a gas lift
reactor (35 L volume) was examined to study the effect of different operating
conditions, i.e. gas recirculation rate, CO feeding rate and addition of bacto-peptone
to the medium (Haddad et al. 2014). The ratio of gas recirculation over CO feeding
rate was the most important parameter affecting the reactor performance, which
kinetically limited both the CO conversion and H2 production rates.
A high amount of biohydrogen (20.5 mM) was produced from 36 mM CO using
an anaerobic microbial consortium in a continuously operated GLR (Sinharoy and
Pakshirajan 2020). The hydrogen production was further enhanced to 30.7 mM with
addition of biologically synthesized iron nanoparticles to the bioreactor. The CO
utilization by the anaerobic consortium was well above 80% for the low CO
concentrations of 7.6 and 15 mM, but gradually depleted with further increase in
the CO concentration.
Immobilized cells of Butyribacterium methylotrophicum have been studied in a
GLR system for syngas fermentation by immobilizing it on different biosupports,
namely celite, molecular sieves, alumina, activated carbon, wood, and ion exchange
resins (Chatterjee et al. 1996). Among the different materials, molecular sieves were
the best suited for cell growth, whereas celite was better for cell attachment than the
other materials. Moreover, the product formation from syngas measured in terms of
total electron content was higher in case of molecular sieves based system.
In addition to GLR operated under mesophilic conditions, its performance has
also been evaluated under thermophilic conditions. For example, hydrogenogenic
CO conversion for biological sulfate reduction was studied using a gas lift reactor
and the effect of different HRT was investigated under thermophilic (55 C) condi-
tions (Sipma et al. 2007). The authors reported that at a high retention time (>5.5 h)
the CO conversion resulted in hydrogenotrophic methane production, whereas H2
production was higher at a short HRT (4 h). Overall, a CO conversion of 85% was
achieved (Sipma et al. 2007).
bioreactor types are effective gas liquid mass transfer and low energy requirement.
Moreover, a MBR provides a high yield, high reaction rate, and increased tolerance
to toxic compounds (e.g. tar, acetylene, NOx and O2) (Campanario and Ortiz 2017).
Besides, this reactor type can operate at high CO partial pressure (PCO). Main
disadvantages are clogging and biofouling of the membrane due to excessive
biomass growth (Munasinghe and Khanal 2010a; Yasin et al. 2015).
Ethanol fermentation from syngas by Clostridium carboxidivorans P7 was stud-
ied in a hollow fibre membrane biofilm reactor (Shen et al. 2014a). The KLa value
(1096.2/h) was higher than most of the commonly used bioreactors. A very high
ethanol concentration of 23.93 g/L was achieved with an ethanol to acetic acid ratio
of 4.79.
CO bioconversion to H2 by Carboxydothermus hydrogenoformans was investi-
gated in a hollow fiber membrane bioreactor under thermophilic (70 C) conditions
(Zhao et al. 2013). The reactor demonstrated consistent performance under different
operational conditions such as liquid flow rate, temperature, CO pressure and CO
loading rate. The CO utilization was 0.44 mol CO/g VSS/day for a partial CO
pressure of 2 atm along with a maximum H2 yield of 92% (mol basis).
A thermophilic reverse membrane bioreactor with a flat plain hydrophilic
polyvinylidene fluoride (PVDF) membrane was used for biomethane production
from CO (Westman et al. 2016). The maximum methane production rate reached
186.0 mL/L/day with a very high syngas consumption, i.e. 7.0, 15.2 and 4.0 mL/L/
day for H2, CO and CO2, respectively. In a separate study, a thermophilic floating
membrane bioreactor was examined for methane production from syngas
(Chandolias et al. 2019). The maximum methane productivity amounted to
34.41 mmol/L/day along with H2 and CO utilization rates of 22 and 50 mmol/L/
day. The results obtained using the membrane bioreactor system for H2 and CO
utilization were, respectively, 38% and 28% higher than the values obtained using a
free cell suspended bioreactor.
The horizontal rotating packed bed (HRPB) reactor is one such novel bioreactor used
for syngas fermentation (Shen et al. 2017). This bioreactor (Fig. 13.3b) is a combi-
nation of two distinct types of bioreactors, viz. the rotating biological contactor and
the packed bed reactor. In this system, similar to the rotating disks in a conventional
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 347
Fig. 13.3 Schematic of some selected novel bioreactor systems for syngas fermentation. (a)
Monolithic biofilm reactor [reused from Shen et al. (2014b), Copyright (2014) with permission
from Elsevier], (b) horizontal rotating packed bed reactor [reused from Shen et al. (2017),
Copyright (2017) with permission from Elsevier], (c) bulk-gas-to-atomized-liquid reactor [reused
from Sathish et al. (2019), Copyright (2019) with permission from Elsevier] and (d) moving bed
biofilm reactor [reused from Sinharoy and Pakshirajan (2021), Copyright (2021) with permission
from Elsevier]
avoided in this bioreactor configuration. This novel bioreactor type was studied for
ethanol production from syngas by Clostridium carboxidivorans P7 strain and a high
ethanol titre and productivity of, respectively, 7.0 g/L and 6.7 g/L/day was achieved
(Shen et al. 2017).
Another novel bioreactor used for enhancing gas-liquid mass transfer during syngas
fermentation is the bulk-gas-to-atomized-liquid (BGAL) contactor reactor (Sathish
et al. 2019). The BGAL reactor consists of a packed bed with biosupport materials to
allow biofilm growth, and syngas saturated liquid micro-droplets are dispersed into
the packing inside the reactor. The syngas-rich liquid droplets then percolate through
the bio-bed and CO is converted to ethanol by the biofilm present on the support
material. The study reported a high ethanol productivity of 746 mg/L/h with an
ethanol to acetic acid ratio of 7.6 from syngas using Clostridium carboxidivorans P7
strain (Sathish et al. 2019).
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 349
In this bioreactor system, liquid is supplied as discrete minute size droplets which
causes a significant increase in the liquid and bulk gas interface (Fig. 13.3c). The
superior mass transfer was demonstrated with oxygen as a model gas, which resulted
in a very high oxygen transfer rate (OTR) of 569 mg/L/min and a KLa value of 2.28/
s. These values are 100 times higher than the values obtained using any other
bioreactor configuration.
Kaldness® biosupport material was used for biofilm formation due to its large
surface area. Use of immobilized biomass in this reactor configuration is more
advantageous than suspended biomass as dispersion of the syngas containing liquid
droplets into the bulk liquid in suspended biomass system (such as CSTR) causes
dilution of the dissolved gases, which in turn reduces the mass transfer and conver-
sion efficiency. The overall energy consumption for transferring the gas to the liquid
phase was also much lower (four-fold) in this BGAL bioreactor in comparison to
a CSTR.
The moving bed biofilm reactor (MBBR; Fig. 13.3d) is another biofilm based reactor
configuration with specially designed biofilm carriers, known as Kaldness®
biosupport material, which offer a large surface area for biofilm formation (Sinharoy
et al. 2020c; Sinharoy and Lens 2020). Although it has not been studied for liquid
biofuel production from syngas, CO bioconversion for H2 production has recently
been reported (Sinharoy and Pakshirajan 2021). A very high H2 production value of
19.5 mmol/L along with ~2 mmol/L of acetate were obtained for 36 mM inlet CO
concentration in the MBBR. The CO utilization was better (>70%) at low inlet CO
concentrations (in the range 9.05–15 mM). Methane production could be completely
avoided in the reactor by 2-Bromoethanosulfonate supplementation to the reactor
during its continuous operation.
The main advantages of the MBBR include high biomass retention, high treat-
ment efficiency, resistance to shock loading conditions and small footprint (Sinharoy
et al. 2019; Sinharoy et al. 2020c). In addition, bed clogging, which is mostly
observed with other attached growth bioreactor configurations, is uncommon in a
MBBR due to continuous mixing. Furthermore, due to the gaseous feed, mixing
inside the reactor can be easily achieved without the need for supplying a separate
(additional) gas stream for mixing, thus keeping the process costs low.
350 A. Sinharoy et al.
In a bioreactor system involving gaseous substrates, e.g. syngas, a high product yield
directly depends upon its bioavailability to the microorganisms, which in turn can be
limited by the high mass transfer rate of the substrate (Liu et al. 2014). A high cell
concentration necessary for the high yield also depends on mass transfer rate of the
gaseous substrates: a low mass transfer rate leads to reduced cell concentrations in
the fermentation media (Yasin et al. 2015). The most important factor limiting the
bioconversion of gaseous substrates is gas-to-liquid mass transfer, similar to oxygen
in aerobic fermentations (Worden et al. 1997). However, limitation due to gas liquid
mass transfer is more problematic for synthesis gas fermentations compared to that
in aerobic fermentations, as the solubility of CO (0.0225 g/kg water) or H2
(0.0015 g/kg water) is less than O2 (0.033 g/kg water) solubility (on a mass basis,
at 30 C) (Abubackar et al. 2011; Munasinghe and Khanal 2012). The different
strategies suggested to improve syngas solubility and gas-liquid mass transfer are
discussed below.
An easy way to improve the gas liquid mass transfer is by enhancing the agitation
speed (rpm) in a bioreactor. Bredwell et al. (1999) reported a high KLa value of
101/h for CO at 450 rpm agitation speed. In another study, an increase in the
agitation speed up to 700 rpm is reported to enhance the CO liquid mass transfer
(Klasson et al. 1993). For achieving a high mechanical agitation speed, a high power
input is required, and, therefore, a high ratio of power required per reactor volume
makes this strategy economically weak for large scale industrial reactors, merely
because of the excessive power costs (Yasin et al. 2015).
A different strategy to improve the mass transfer is by employing different
impeller design (Ungerman and Heindel 2007). The effect of six different impeller
configurations, namely Rushton-type, Philadelphia Mixing concave (hollow blade)
turbine, Philadelphia Mixing pitched blade turbine (PBT), Philadelphia Mixing LS
hydrofoil, Lightnin A315 fluidfoil and Lightnin A310, on the gas liquid mass
transfer rate was analyzed. The Rushton-type performed better compared to the
other five impeller types. Also, the dual Rushton type showed a 27% higher gas
liquid mass transfer compared to the conventional single Rushton type impeller.
However, its performance in terms of the ratio of volumetric mass transfer coeffi-
cient to power input was poor. Hence, it should be emphasized that improved gas
liquid mass transfer in a stirred system demands a high power consumption and cost
associated with it. Also, the increase in the agitation speed to enhance the gas liquid
mass transfer may be damaging to shear sensitive microorganisms, thereby hamper-
ing microbial growth in the bioreactor (Munasinghe and Khanal 2012).
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 351
13.5.2 Additives
Fig. 13.4 Schematic of nanoparticle mediated enhancement of the gas-liquid mass transfer
References
Abubackar HN, Veiga MC, Kennes C (2011) Biological conversion of carbon monoxide: rich
syngas or waste gases to bioethanol. Biofuels Bioprod Biorefin 5(1):93–114
Abubackar HN, Veiga MC, Kennes C (2015) Carbon monoxide fermentation to ethanol by
Clostridium autoethanogenum in a bioreactor with no accumulation of acetic acid. Bioresour
Technol 186:122–127
Abubackar HN, Fernández-Naveira Á, Veiga MC, Kennes C (2016) Impact of cyclic pH shifts on
carbon monoxide fermentation to ethanol by Clostridium autoethanogenum. Fuel 178:56–62
Abubackar HN, Veiga MC, Kennes C (2018) Production of acids and alcohols from syngas in a
two-stage continuous fermentation process. Bioresour Technol 253:227–234
Aghbashlo M, Tabatabaei M, Dadak A, Younesi H, Najafpour G (2016) Exergy-based performance
analysis of a continuous stirred bioreactor for ethanol and acetate fermentation from syngas via
Wood–Ljungdahl pathway. Chem Eng Sci 143:36–46
Ahmad AA, Zawawi NA, Kasim FH, Inayat A, Khasri A (2016) Assessing the gasification
performance of biomass: a review on biomass gasification process conditions, optimization
and economic evaluation. Renew Sustain Energy Rev 53:1333–1347
Alves JI, van Gelder AH, Alves MM, Sousa DZ, Plugge CM (2013) Moorella stamsii sp. nov., a
new anaerobic thermophilic hydrogenogenic carboxydotroph isolated from digester sludge. Int J
Syst Evol Microbiol 63 (11):4072–4076
Amos WA (2004) Biological water-gas shift conversion of carbon monoxide to hydrogen: mile-
stone completion report (No. NREL/MP-560-35592). National Renewable Energy Lab.,
Golden, CO
Asimakopoulos K, Gavala HN, Skiadas IV (2018) Reactor systems for syngas fermentation
processes: a review. Chem Eng J 348:732–744
Asimakopoulos K, Łężyk M, Grimalt-Alemany A, Melas A, Wen Z, Gavala HN, Skiadas IV (2020)
Temperature effects on syngas biomethanation performed in a trickle bed reactor. Chem Eng J
393:124739
Asimakopoulos K, Grimalt-Alemany A, Lundholm-Høffner C, Gavala HN, Skiadas IV (2021a)
Carbon sequestration through syngas biomethanation coupled with H2 supply for a clean
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 353
Cowger JP, Klasson KT, Ackerson MD, Clausen E, Caddy JL (1992) Mass-transfer and kinetic
aspects in continuous bioreactors using Rhodospirillum rubrum. Appl Biochem Biotechnol
34(1):613–624
Datar RP, Shenkman RM, Cateni BG, Huhnke RL, Lewis RS (2004) Fermentation of biomass-
generated producer gas to ethanol. Biotechnol Bioeng 86(5):587–594
Devarapalli M, Atiyeh HK, Phillips JR, Lewis RS, Huhnke RL (2016) Ethanol production during
semi-continuous syngas fermentation in a trickle bed reactor using Clostridium ragsdalei.
Bioresour Technol 209:56–65
Diender M, Stams AJ, Sousa DZ (2015) Pathways and bioenergetics of anaerobic carbon monoxide
fermentation. Front microbiol 6:1275
Diender M, Uhl PS, Bitter JH, Stams AJ, Sousa DZ (2018) High rate biomethanation of carbon
monoxide-rich gases via a thermophilic synthetic coculture. ACS Sustain Chem Eng
6(2):2169–2176
Do YS, Smeenk J, Broer KM, Kisting CJ, Brown R, Heindel TJ, Bobik TA, DiSpirito AA (2007)
Growth of Rhodospirillum rubrum on synthesis gas: conversion of CO to H2 and
poly-β-hydroxyalkanoate. Biotechnol Bioeng 97(2):279–286
Ebrahimi S, Kleerebezem R, Kreutzer MT, Kapteijn F, Moulijn JA, Heijnen JJ, Van Loosdrecht
MCM (2006) Potential application of monolith packed columns as bioreactors, control of
biofilm formation. Biotechnol Bioeng 93(2):238–245
Ferry JG (2010) CO in methanogenesis. Ann Microbiol 60(1):1–12
Griffin DW, Schultz MA (2012) Fuel and chemical products from biomass syngas: a comparison of
gas fermentation to thermochemical conversion routes. Environ Prog Sustain Energy
31(2):219–224
Grimalt-Alemany A, Skiadas IV, Gavala HN (2018) Syngas biomethanation: state-of-the-art review
and perspectives. Biofuel Bioprod Biorefin 12(1):139–158
Grimalt-Alemany A, Asimakopoulos K, Skiadas IV, Gavala HN (2020) Modeling of syngas
biomethanation and catabolic route control in mesophilic and thermophilic mixed microbial
consortia. Appl Energy 262:114502
Guiot SR, Cimpoia R, Carayon G (2011) Potential of wastewater-treating anaerobic granules for
biomethanation of synthesis gas. Environ Sci Technol 45(5):2006–2012
Gunes B (2021) A critical review on biofilm-based reactor systems for enhanced syngas fermen-
tation processes. Renew Sustain Energy Rev 143:110950
Haddad M, Cimpoia R, Guiot SR (2014) Performance of Carboxydothermus hydrogenoformans in
a gas-lift reactor for syngas upgrading into hydrogen. Int J Hydrogen Energy 39(6):2543–2548
Hassan SS, Williams GA, Jaiswal AK (2019) Moving towards the second generation of lignocel-
lulosic biorefineries in the EU: drivers, challenges, and opportunities. Renew Sustain Energy
Rev 101:590–599
Henstra AM, Stams AJ (2011) Deep conversion of carbon monoxide to hydrogen and formation of
acetate by the anaerobic thermophile Carboxydothermus hydrogenoformans. Int J Microbiol.
https://doi.org/10.1155/2011/641582
Henstra AM, Sipma J, Rinzema A, Stams AJ (2007) Microbiology of synthesis gas fermentation for
biofuel production. Curr Opin Biotechnol 18(3):200–206
Hurst KM, Lewis RS (2010) Carbon monoxide partial pressure effects on the metabolic process of
syngas fermentation. Biochem Eng J 48(2):159–165
Ismail KSK, Najafpour G, Younesi H, Mohamed AR, Kamaruddin AH (2008) Biological hydrogen
production from CO: bioreactor performance. Biochem Eng J 39(3):468–477
Isom CE, Nanny MA, Tanner RS (2015) Improved conversion efficiencies for n-fatty acid reduction
to primary alcohols by the solventogenic acetogen “Clostridium ragsdalei”. J Ind Microbiol
Biotechnol 42(1):29–38
Jing Y, Campanaro S, Kougias P, Treu L, Angelidaki I, Zhang S, Luo G (2017) Anaerobic granular
sludge for simultaneous biomethanation of synthetic wastewater and CO with focus on the
identification of CO-converting microorganisms. Water Res 126:19–28
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 355
Johnsson F, Kjärstad J, Rootzén J (2019) The threat to climate change mitigation posed by the
abundance of fossil fuels. Clim Policy 19(2):258–274
Jung GY, Kim JR, Park JY, Park S (2002) Hydrogen production by a new chemoheterotrophic
bacterium Citrobacter sp. Y19. Int J Hydrogen Energy 27(6):601–610
Jurasz J, Canales FA, Kies A, Guezgouz M, Beluco A (2020) A review on the complementarity of
renewable energy sources: concept, metrics, application and future research directions. Sol
Energy 195:703–724
Kapic A, Jones ST, Heindel TJ (2006) Carbon monoxide mass transfer in a syngas mixture. Ind Eng
Chem Res 45(26):9150–9155
Kiefer D, Merkel M, Lilge L, Henkel M, Hausmann R (2021) From acetate to bio-based products:
underexploited potential for industrial biotechnology. Trends Biotechnol 39(4):397–411
Klasson KT, Gupta A, Clausen EC, Gaddy JL (1993) Evaluation of mass-transfer and kinetic
parameters for Rhodospirillum rubrum in a continuous stirred tank reactor. Appl Biochem
Biotechnol 39(1):549–557
Kumar M, Sinharoy A, Pakshirajan K (2018) Process integration for biological sulfate reduction in
a carbon monoxide fed packed bed reactor. J Environ Manag 219:294–303
Kundiyana DK, Huhnke RL, Wilkins MR (2010) Syngas fermentation in a 100-L pilot scale
fermentor: design and process considerations. J Biosci Bioeng 109(5):492–498
Kundiyana DK, Huhnke RL, Wilkins MR (2011) Effect of nutrient limitation and two-stage
continuous fermentor design on productivities during “Clostridium ragsdalei” syngas fermen-
tation. Bioresour Technol 102(10):6058–6064
Lagoa-Costa B, Abubackar HN, Fernández-Romasanta M, Kennes C, Veiga MC (2017) Integrated
bioconversion of syngas into bioethanol and biopolymers. Bioresour Technol 239:244–249
Lee PH, Ni SQ, Chang SY, Sung S, Kim SH (2012) Enhancement of carbon monoxide mass
transfer using an innovative external hollow fiber membrane (HFM) diffuser for syngas
fermentation: experimental studies and model development. Chem Eng J 184:268–277
Li Y, Wang Z, He Z, Luo S, Su D, Jiang H, Zhou H, Xu Q (2020) Effects of temperature, hydrogen/
carbon monoxide ratio and trace element addition on methane production performance from
syngas biomethanation. Bioresour Technol 295:122296
Liu K, Atiyeh HK, Tanner RS, Wilkins MR, Huhnke RL (2012) Fermentative production of ethanol
from syngas using novel moderately alkaliphilic strains of Alkalibaculum bacchi. Bioresour
Technol 104:336–341
Liu K, Atiyeh HK, Stevenson BS, Tanner RS, Wilkins MR, Huhnke RL (2014) Continuous syngas
fermentation for the production of ethanol, n-propanol and n-butanol. Bioresour Technol 151:
69–77
Liu Y, Wan J, Han S, Zhang S, Luo G (2016) Selective conversion of carbon monoxide to hydrogen
by anaerobic mixed culture. Bioresour Technol 202:1–7
Maddipati P, Atiyeh HK, Bellmer DD, Huhnke RL (2011) Ethanol production from syngas by
Clostridium strain P11 using corn steep liquor as a nutrient replacement to yeast extract.
Bioresour Technol 102(11):6494–6501
Martin DR, Lundie LL, Kellum R, Drake HL (1983) Carbon monoxide-dependent evolution of
hydrogen by the homoacetate-fermenting bacterium Clostridium thermoaceticum. Curr
Microbiol 8(6):337–340
Mohammadi M, Najafpour GD, Younesi H, Lahijani P, Uzir MH, Mohamed AR (2011) Biocon-
version of synthesis gas to second generation biofuels: a review. Renew Sustain Energy Rev
15(9):4255–4273
Molino A, Chianese S, Musmarra D (2016) Biomass gasification technology: the state of the art
overview. J Energy Chem 25(1):10–25
Mörsdorf G, Frunzke K, Gadkari D, Meyer O (1992) Microbial growth on carbon monoxide.
Biodegradation 3(1):61–82
Munasinghe PC, Khanal SK (2010a) Biomass-derived syngas fermentation into biofuels: opportu-
nities and challenges. Bioresour Technol 101(13):5013–5022
356 A. Sinharoy et al.
Munasinghe PC, Khanal SK (2010b) Syngas fermentation to biofuel: evaluation of carbon mon-
oxide mass transfer coefficient (kLa) in different reactor configurations. Biotechnol Prog
26(6):1616–1621
Munasinghe PC, Khanal SK (2012) Syngas fermentation to biofuel: evaluation of carbon monoxide
mass transfer and analytical modeling using a composite hollow fiber (CHF) membrane
bioreactor. Bioresour Technol 122:130–136
Munasinghe PC, Khanal SK (2014) Evaluation of hydrogen and carbon monoxide mass transfer
and a correlation between the myoglobin-protein bioassay and gas chromatography method for
carbon monoxide determination. RSC Adv 4(71):37575–37581
Najafpour G, Younesi H, Mohamed AR (2003) Continuous hydrogen production via fermentation
of synthesis gas. Pet Coal 45(3/4):154–158
Navarro SS, Cimpoia R, Bruant G, Guiot SR (2014) Specific inhibitors for identifying pathways for
methane production from carbon monoxide by a nonadapted anaerobic mixed culture. Can J
Microbiol 60(6):407–415
Negi BB, Sinharoy A, Pakshirajan K (2020) Selenite removal from wastewater using fungal
pelleted airlift bioreactor. Environ Sci Pollut Res 27(1):992–1003
Oelgeschläger E, Rother M (2008) Carbon monoxide-dependent energy metabolism in anaerobic
bacteria and archaea. Arch Microbiol 190(3):257–269
Oelgeschläger E, Rother M (2009) Influence of carbon monoxide on metabolite formation in
Methanosarcina acetivorans. FEMS Microbiol Lett 292(2):254–260
Oh YK, Kim YJ, Park JY, Lee TH, Kim MS, Park S (2005) Biohydrogen production from carbon
monoxide and water by Rhodopseudomonas palustris P4. Biotechnol Bioprocess Eng 10(3):270
Pakpour F, Najafpour G, Tabatabaei M, Tohidfar M, Younesi H (2014) Biohydrogen production
from CO-rich syngas via a locally isolated Rhodopseudomonas palustris PT. Bioprocess
Biosyst Eng 37(5):923–930
Pakshirajan K, Mal J (2013) Biohydrogen production using native carbon monoxide converting
anaerobic microbial consortium predominantly Petrobacter sp. Int J Hydrogen Energy
38(36):16020–16028
Park S, Yasin M, Kim D, Park HD, Kang CM, Kim DJ, Chang IS (2013) Rapid enrichment of
(homo) acetogenic consortia from animal feces using a high mass-transfer gas-lift reactor fed
with syngas. J Ind Microbiol Biotechnol 40(9):995–1003
Rajagopalan S, Datar RP, Lewis RS (2002) Formation of ethanol from carbon monoxide via a new
microbial catalyst. Biomass Bioenergy 23(6):487–493
Riegler P, Chrusciel T, Mayer A, Doll K, Weuster-Botz D (2019) Reversible retrofitting of a stirred-
tank bioreactor for gas-lift operation to perform synthesis gas fermentation studies. Biochem
Eng J 141:89–101
Riggs SS, Heindel TJ (2006) Measuring carbon monoxide gas—liquid mass transfer in a stirred
tank reactor for syngas fermentation. Biotechnol Prog 22(3):903–906
Roy AS, Chingkheihunba A, Pakshirajan K (2016) An overview of production, properties, and uses
of biodiesel from vegetable oil. Green Fuels Technol 83–105
Sancho Navarro S, Cimpoia R, Bruant G, Guiot SR (2016) Biomethanation of syngas using
anaerobic sludge: shift in the catabolic routes with the CO partial pressure increase. Front
Microbiol 7:1188
Sathish A, Sharma A, Gable P, Skiadas I, Brown R, Wen Z (2019) A novel bulk-gas-to-atomized-
liquid reactor for enhanced mass transfer efficiency and its application to syngas fermentation.
Chem Eng J 370:60–70
Schuchmann K, Müller V (2016) Energetics and application of heterotrophy in acetogenic bacteria.
Appl Environ Microbiol 82(14):4056–4069
Shen Y, Brown R, Wen Z (2014a) Syngas fermentation of Clostridium carboxidivoran P7 in a
hollow fiber membrane biofilm reactor: evaluating the mass transfer coefficient and ethanol
production performance. Biochem Eng J 85:21–29
Shen Y, Brown R, Wen Z (2014b) Enhancing mass transfer and ethanol production in syngas
fermentation of Clostridium carboxidivorans P7 through a monolithic biofilm reactor. Appl
Energy 136:68–76
13 Syngas Fermentation for Bioenergy Production: Advances in Bioreactor Systems 357
Stoll IK, Boukis N, Sauer J (2020) Syngas fermentation to alcohols: reactor technology and
application perspective. Chem Ing Tech 92(1-2):125–136
Strübing D, Huber B, Lebuhn M, Drewes JE, Koch K (2017) High performance biological
methanation in a thermophilic anaerobic trickle bed reactor. Bioresour Technol 245:1176–1183
Strübing D, Moeller AB, Mößnang B, Lebuhn M, Drewes JE, Koch K (2018) Anaerobic thermo-
philic trickle bed reactor as a promising technology for flexible and demand-oriented H2/CO2
biomethanation. Appl Energy 232:543–554
Subramani V, Gangwal SK (2008) A review of recent literature to search for an efficient catalytic
process for the conversion of syngas to ethanol. Energy Fuels 22(2):814–839
Swart SD (2013) Design, modelling and construction of a scalable dual fluidised bed reactor for the
pyrolysis of biomass. Doctoral dissertation, University of Pretoria
Ungerman AJ, Heindel TJ (2007) Carbon monoxide mass transfer for syngas fermentation in a
stirred tank reactor with dual impeller configurations. Biotechnol Prog 23(3):613–620
Vanzin GF, Huang J, Smolinski S, Kronoveter K, Maness PC (2002) Biological hydrogen from fuel
gases. In: Proceedings of the US DOE Hydrogen Program Review
Wainaina S, Horváth IS, Taherzadeh MJ (2018) Biochemicals from food waste and recalcitrant
biomass via syngas fermentation: a review. Bioresour Technol 248:113–121
Wang W, Xie L, Luo G, Zhou Q, Angelidaki I (2013) Performance and microbial community
analysis of the anaerobic reactor with coke oven gas biomethanation and in situ biogas
upgrading. Bioresour Technol 146:234–239
Weghoff MC, Müller V (2016) CO metabolism in the thermophilic acetogen Thermoanaerobacter
kivui. Appl Environ Microbiol 82(8):2312–2319
Westman SY, Chandolias K, Taherzadeh MJ (2016) Syngas biomethanation in a semi-continuous
reverse membrane bioreactor (RMBR). Fermentation 2(2):8
Wolfrum EJ, Watt AS (2001) Bioreactor design studies for a novel hydrogenproducing
bacterium. In: Proceedings of the 2001 DOE Hydrogen Program Review
Worden RM, Bredwell MD, Grethlein AJ (1997) Engineering issues in synthesis-gas fermentations.
ACS Symp Ser Am Chem Soc 18:320–335
Yasin M, Jeong Y, Park S, Jeong J, Lee EY, Lovitt RW, Kim BH, Lee J, Chang IS (2015) Microbial
synthesis gas utilization and ways to resolve kinetic and mass-transfer limitations. Bioresour
Technol 177:361–374
Yoneda Y, Yoshida T, Kawaichi S, Daifuku T, Takabe K, Sako Y (2012) Carboxydothermus
pertinax sp. nov., a thermophilic, hydrogenogenic, Fe (III)-reducing, sulfur-reducing
carboxydotrophic bacterium from an acidic hot spring. Int J Syst Evol Microbiol 62
(Pt_7)):1692–1697
Younesi H, Najafpour G, Ismail KSK, Mohamed AR, Kamaruddin AH (2008) Biohydrogen
production in a continuous stirred tank bioreactor from synthesis gas by anaerobic photosyn-
thetic bacterium: Rhodopirillum rubrum. Bioresour Technol 99(7):2612–2619
Zavarzina DG, Sokolova TG, Tourova TP, Chernyh NA, Kostrikina NA, Bonch-Osmolovskaya EA
(2007) Thermincola ferriacetica sp. nov., a new anaerobic, thermophilic, facultatively
chemolithoautotrophic bacterium capable of dissimilatory Fe (III) reduction. Extremophiles
11(1):1–7
Zhao Y, Haddad M, Cimpoia R, Liu Z, Guiot SR (2013) Performance of a Carboxydothermus
hydrogenoformans-immobilizing membrane reactor for syngas upgrading into hydrogen. Int J
Hydrogen Energy 38(5):2167–2175
Zheng X, Ying Z, Wang B, Chen C (2018) Hydrogen and syngas production from municipal solid
waste (MSW) gasification via reusing CO2. Appl Therm Eng 144:242–247
Zhu H, Shanks BH, Heindel TJ (2008) Enhancing CO water mass transfer by functionalized
MCM41 nanoparticles. Ind Eng Chem Res 47(20):7881–7887
Zhu H, Shanks BH, Heindel TJ (2009) Effect of electrolytes on CO water mass transfer. Ind Eng
Chem Res 48(6):3206–3210
Zuidervaart E, Reuter MA, Heerema RH, Van der Lans RGJM, Derksen JJ (2000) Effect of
dissolved metal sulphates on gas-liquid oxygen transfer in agitated quartz and pyrite slurries.
Miner Eng 13(14–15):1555–1564
Part VI
Life Cycle Analysis
Chapter 14
Up and Downstream Technologies
of Anaerobic Digestion from Life Cycle
Assessment Perspective
Seyedeh Nashmin Elyasi, Hadis Marami, Li He, Ali Kaab, Junting Pan,
Hongbin Liu, and Benyamin Khoshnevisan
Abstract The development of anaerobic digestion (AD) plants has been considered
as a solution to overcome the depletion of fossil resources and the increasing
environmental pollution caused by the over-use of fossil fuels. However, there are
some challenges which may undermine the sustainability of the AD including its
associated up- and downstream strategies. In order to sustain the development of AD
plants it is necessary to calculate the environmental hotspots through their whole life
cycle and represent the possible solutions to mitigate the potential pollution. Among
different methods, life cycle assessment (LCA) is widely employed to quantify and
evaluate the environmental impacts of AD and its related up- and downstream
technologies. This chapter comprehensively summarizes the environmental impacts
of AD plants from an LCA point of view and proposes advanced strategies to
alleviate the negative impacts and bring biogas plants into sustainable circular
bio-economy.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 361
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_14
362 S. N. Elyasi et al.
14.1 Introduction
In the last decades, the development of anaerobic digestion (AD) plants have
contributed to achieving both environmental pollution mitigation and energy inde-
pendency by utilizing waste streams, such as livestock manure and organic fraction
of municipal solid waste (OFMSW) to produce biogas and biofertilizer (Duan et al.
2020). Biogas and nutrient-rich digestate produced by the AD plants bring about
significant environmental savings, specifically, in terms of mitigation of greenhouse
gas (GHG) emissions (Fusi et al. 2016). However, properly designed AD plants
consist of upstream and downstream technologies for the appropriate treatment of
the feedstock and suitable biogas and digestate conditioning. Some upstream and
downstream processes rely on the consumption of considerable amounts of
chemicals as well as energy; thereby they could significantly undermine the envi-
ronmental sustainability of the AD plants (Khoshnevisan et al. 2016, 2018b;
Tsapekos et al. 2021). Furthermore, taking the environmental savings into consid-
eration, biogas utilization and digestate management are leading criteria to assess the
sustainability of AD facilities (Ardolino et al. 2020; Kohlheb et al. 2020). Therefore,
to sustain the processing of any type of feedstock in AD plants, the environmental
footprints of every constitution of the AD processes should be concerned.
In this regard, life cycle assessment (LCA), covering energy flow, mass flow,
GHG emissions, and other environmental impacts, is a well-established method
which can be employed as a standard tool to evaluate the sustainability of any
given system through its whole life cycle (Finkbeiner 2014). LCA enables the
quantification of the environmental impacts of the systems under consideration
and find the environmental hotspots and possible opportunities for future improve-
ments (Dastjerdi et al. 2021). In this context, considerable studies have already been
published under the framework of LCA to assess the environmental impacts of the
AD plants and their related subsystems. This chapter aims to comprehensively
review the state-of-the-art progress in the sustainability of AD plants by scrutinizing
the environmental impacts caused by both up and downstream processes (i.e.,
pretreatment technologies, biogas desulfurization and upgrading, as well as biogas
and digestate utilization) to show the recent advancements and the roadmaps for
future work.
LCA has long been used as a valuable tool to assess the environmental footprints of
AD systems. Since AD has been considered as a waste management strategy along
with a decentralized energy technology, LCA has been widely employed during the
past decades to scrutinize the sustainability of AD plants and their associated
sub-systems as part of the waste management system and waste to energy scenario
(De Jong et al. 2017). The AD technology has been in use for over 100 years and is
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 363
Organic waste, agricultural waste, manure, and energy crops are the most frequently
used biomass in AD plants (Aziz et al. 2019). The environmental performance of AD
systems largely depends on the feedstock used within the plants and the amount of
biogas exploited (Montgomery and Bochmann 2014). However, some specific types
of biomass such as lignocellulosic materials are critically slow to be degraded during
the AD process to produce reasonable amount of biogas (Lee et al. 2020). Accord-
ingly, prior to being fed into anaerobic digesters, specific pretreatment technologies
are needed depending on the feedstock used and the AD technology (Carrere et al.
2016).
AD consists of four main steps, i.e., hydrolysis (formation of sugar, long chain
fatty acid, amino acid), acidogenesis (acid formation i.e., alcohols and carbonic
acid), acetogenesis (acetic acid formation, i.e., CH3OH, NH3, H2S, and CO2), and
methanogenesis (biogas) (Koupaie et al. 2019). Pretreatments target enhancing the
hydrolysis step to facilitate the breakdown of the complex molecules, shortening the
biodegradation time, and enhancing biomethane formation (Tabatabaei et al. 2020).
Pretreatment technologies are divided into physical, chemical, biological, and com-
bined process (Gnaoui et al. 2020). Further information and subcategories of main
classes of pretreatment methods are illustrated in Fig. 14.1.
From the LCA point of view, any pretreatment technology performed upstream of
AD, apart from the effectiveness for increasing the biomethane yield, should be
Table 14.1 Overview of studies related to LCA of AD systems
364
System Boundary
Feedstock
production
Surveyed and Biogas Product Impact assessment
location Goal and Scope collection AD Upgrading use method Feedstock type Reference
Italy Co-digestion of sewage No Yes No Yes ILCD Metrics Co-digested fruit and veg- Di Maria et al.
sludge and fruit and vege- etable waste with waste (2016)
table wastes mixed sludge from
municipal wastewater
treatment plants
USA Compare areal productiv- No Yes No Yes No Dairy waste water Ciliberti et al.
ities of AD and wind farms (2016)
France LCA of microalgae AD Yes Yes Yes Yes ReCiPe 2008 Microalgae Shimako et al.
(2016)
Spain AD of sewage sludge No Yes No Yes ReCiPe v.1.08 Sewage sludge Blanco et al.
(midpoint) (2016)
Sweden AD of food waste and No Yes No Yes CML 2001 (GHG Municipal solid waste Eriksson et al.
sewage sludge results only) (2016)
Germany Electricity production Yes Yes Yes Yes ReCiPe v.1.06 Co-digestion of maize Junne and
from agriculture feedstock (midpoint) and CED (or sunflower, horse Junne (2016)
v.1.08 manure and biogenic
waste), grass, rye silage,
and chicken manure
Germany Compare organic WM Yes No No No A method with sim- Organic waste Jensen et al.
systems ilar indicators as (2016)
specified in ReCiPe
midpoint metrics
Austria Biogas from pretreated Yes No Yes No ReCiPe v.1.08 Maize silage Kral et al.
maize silage (mid-point) and (2016)
ILCD 2011
S. N. Elyasi et al.
14
USA Comparative analysis of No Yes No No IMPACT 2002+ and Wastewater Pericault et al.
TRACI 2.0 (2017)
Turkey AD of agricultural farm No Yes No Yes EDIP 2003 Agricultural wastes Nayal et al.
waste (2016)
Australia Co-digestion of sewage No Yes Yes Yes CML-IA Sewage sludge and sorted Edwards et al.
sludge and municipal food food waste (2017)
waste
Germany Electricity production Yes Yes Yes Yes ReCiPe v.1.06 Co-digestion of marine Ertem et al.
from microalgal feedstock (midpoint) Microalgae or maize, (2016)
grass, rye silage, and
chicken manure
France AD of sewage sludge and No Yes Yes Yes GHG Emissions; Sewage sludge Collet et al.
methanation ReCiPe v.1.08 (2017)
(endpoint)
Denmark Different strategies for No Yes Yes No IMPACT 2002+ Household waste Khoshnevisan
energy and nutrient recov- et al. (2018b)
ery from source sorted
organic fraction of AD
Denmark Environmental impacts of Yes Yes Yes Yes IMPACT 2002+ Grass Tsapekos et al.
biogas production from (2021)
grass
Denmark Biowaste biorefinery No Yes Yes Yes IMPACT 2002+ Municipal solid waste Khoshnevisan
concepts et al. (2020b)
China LCA of AD of pig manure No Yes No Yes IMPACT 2002+ Pig manure Duan et al.
(2020)
Adapted from Rajendran and Murthy (2019)
Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . .
365
366
Fig. 14.1 Pretreatment technologies utilized in AD plants. LS, lignocellulosic. Characteristics taken from Chiu and Lo (2016), Shrestha et al. (2020), Torres and
Lloréns (2008), Wang et al. (2015) and Zhang et al. (2011)
S. N. Elyasi et al.
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 367
eco-friendly in terms of water, energy, and chemical consumption with less envi-
ronmental burdens (Ravindran et al. 2018). Table 14.2 summarizes the previous
studies with focus on pretreatments prior to AD. Among the existing technologies
for the pretreatment of bio-feedstock, biological routes have low energy demand and
could take part in the AD plants at low temperature in the absence of chemical
additives (Chiu and Lo 2016; Tabatabaei et al. 2020). The enzymatic pretreatment
has shown high efficiency to significantly improve the biogas production potential
while being independent of water consumption and has minor environmental
impacts (Asgher et al. 2014). Besides, fungi, specifically white-rot fungi, enable
removal of the environmental pollutants from solid and liquid waste before being
used for AD (Jayachandra et al. 2011; Montgomery and Bochmann 2014).
Biological pretreatment technologies, such as micro-aeration and anaerobic
methods, are carried out too slowly compared with physical and chemical processes
(Bochmann and Montgomery 2013). Therefore, some of the organic matter is
degraded to CO2 instead of being converted into methane in the subsequent AD
process (Chiu and Lo 2016). In order to alleviate this bottleneck to some extent, it is
recommended to combine microaeration with an anaerobic pretreatment reactor and
thereby enhancing the methane yield (Montgomery and Bochmann 2014). Apart
from the mentioned advantages and disadvantages, biological methods are com-
monly used in combination with other physical-chemical technologies to tackle the
aforementioned limitations (Khoshnevisan et al. 2018b; Tabatabaei et al. 2020).
Accordingly, the requirements for auxiliary pretreatment technologies have
undermined the wide industrial application of biological methods in AD systems
(Tabatabaei et al. 2020).
Microalgae Thermal 1 kWh Comparing the hydrothermal and GHG Xiao et al.
biogas solar (2020)
driven hydrothermal
GWP, global warming potential; TP, toxicity potential; EP, eutrophication potential; AP, acidification potential; GHG, greenhouse gas emissions; ODP, ozone
depletion potential; PMF, particulate matter formation; OFP, ozone formation potential; POCP, photo ozone creation potential; ADP, abiotic depletion; POF,
photochemical oxidant formation; TA, terrestrial acidification; ME, marine eutrophication; SFD, fossil resource depletion; HT, human toxicity; SSOHW, source
sorted organic fraction of household waste; OFMSW, organic fraction of municipal solid waste
Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . .
369
370 S. N. Elyasi et al.
sorting of food waste into a substrate and reject fraction as well as producing a
homogeneous feedstock for the AD process to have a higher methane recovery
(Khoshnevisan et al. 2018b). Furthermore, thermal pretreatment methods which falls
into physical pretreatment technology accelerates the hydrolysis of biofeedstock by
disintegrating cell membranes and facilitating the dissolution of recalcitrant organic
compounds, specially sewage sludge, microalgae, and food waste (Chen et al. 2012;
Gnaoui et al. 2020; Karuppiah and Azariah 2019). This method is normally
performed at very high temperature and thus requires a significant amount of heat
(Rajput et al. 2018). Although thermal pretreatment technology increases the meth-
ane yield of specific feedstock which otherwise have a low biodegradability during
the AD process, it may not increase the net environmental savings of AD systems
due to the high energy demand (Mills et al. 2014). It is worth mentioning that, if the
heat demand of thermal pretreatment methods could be supplied from renewable
sources such as solar systems, the environmental performance would be more
reasonable than using fossil-based heat sources. To illustrate, the total GHG emis-
sions from the production of biogas with hydrothermal pretreatment and solar-driven
hydrothermal pretreatment methods were estimated at –129.94, and –166.13 g CO2
eq/kWh biogas, respectively (Xiao et al. 2020).
Last but not least, ultrasound and electrokinetic disintegration are available
physical pretreatment technologies that are not applied for lignocellulosic substrate
and are only specified for sewage sludge (Dhar et al. 2012; Montgomery and
Bochmann 2014). Previous studies show that ultrasound and electrokinetic disinte-
gration could be useful technologies to improve the AD fed sludge by disordering
microbial cell walls in the liquid (Bochmann and Montgomery 2013; Cano et al.
2015; Cartes et al. 2018). However, generally, these pretreatment methods consume
electricity and reduce the net environmental profits since a part of exploited elec-
tricity from biogas should be utilized internally (Cano et al. 2015).
Chemical pretreatment approaches are popular for treating substrate with a high
share of lignocellulosic composition in which various types of chemicals such as
HNO3, H2SO4, HCl, H3PO4, KOH, NaOH, and CaO are applied to break the
complex structures and reduce the digester hydraulic retention time (Vannarath
and Thalla 2019; Zhen et al. 2017). The application of chemical pretreatment routes
could increase the methane yield up to 200% (Torres and Lloréns 2008; Wang et al.
2015; Zhang et al. 2011). For example, oxidative pretreatment by utilizing hydrogen
peroxide doubled the biogas production from rice straw (Song et al. 2012). However,
the use of chemicals for this purpose, specifically in the case of alkali and acid
methods, increases the concerns over the sustainability of chemical pretreatment
(Smullen et al. 2017, 2019). Accordingly, the chemical methods have not been
widely used in large scale biogas plants despite their vast utilization for the produc-
tion of ethanol (Carrere et al. 2016). Nevertheless, there are increasing attempts at
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 371
In order to take the advantage of previously mentioned technologies and reduce their
bottlenecks to some extent, combined processes have been developed which cannot
be classified into the former categories (i.e., mechanical, thermal or chemical
pretreatment). Such pretreatment methods are more efficient since they are a com-
bination of two or more types of main pretreatment technologies (Montgomery and
Bochmann 2014). These methods are in forms of physicochemical, physical-
thermal, thermo-chemical, mechanical-chemical, and bio-chemical technologies.
Such integrated technologies require less energy, lower capital costs, and easy
operation compared to each individual technology (Shrestha et al. 2020). As a matter
of fact, by combining chemical methods with mechanical, thermal, and biological
technologies, the amounts of energy and chemical consumption by individual
technologies are reduced while the biomethane yield is increased. For example,
thermal-alkaline and microwave-alkaline pre-treatment have been shown to decrease
the consumption of alkali compounds whilst increasing the biogas production (Pilli
et al. 2020). As another example, some integrated thermal-based technologies such
as steam explosion and extrusion consume lower energy compared to individual
thermal and hydrothermal methods. Nonetheless, they are carried out at elevated
temperature with high positive effect on the methane yield on straw in AD system
(Martínez-Gutiérrez 2018).
In addition to the AD and its upstream processes, downstream technologies also play
a critical role in the sustainability of AD plants. Downstream processes include
biogas conditioning (i.e., purification and upgrading), biogas utilization, and
digestate management. In the following sub-sections, the effects of downstream
processes and technologies on the sustainability of AD plants are discussed in detail.
The generated biogas in the AD process consists of 40–70% (v/v) CH4, 30–60%
(v/v) CO2, 0–1% (v/v) H2, 0–3% (v/v) H2S, and 0–5% (v/v) other gases (Tabatabaei
372 S. N. Elyasi et al.
Fig. 14.2 Physiochemical and biological desulfurization technologies from LCA perspective
and Ghanavati 2018). For further application of the biogas and the subsequent
production of heat and electricity in conventional conversion systems including
large turbines, gas micro-turbines, combined heat and power (CHP), and internal
combustion engines, the H2S content of biogas should be removed to avoid corro-
sions (Santos-Clotas et al. 2019). The removal of H2S, which is formed by microbial
reduction of sulfur compounds, is the most important downstream process for biogas
conditioning since the H2S content of biogas can cause corrosion or toxicity (Hauser
2017). There are various physicochemical and biological technologies available for
H2S removal. Figure 14.2 illustrates all types of available technologies employed for
H2S removal. Previous studies have widely characterized various types of desulfur-
ization technologies (Xiao et al. 2017; Zăbavă et al. 2019; Zulkefli et al. 2016).
Taking into account the desulfurization of biogas from LCA point of view, in all
the evaluated technologies so far, daily operation of desulfurization is the highest
contributor to the overall environmental impacts rather than the construction and
disposal of desulfurization units alongside their auxiliary equipment (Cano et al.
2018). Among various technologies commercialized, bio-desulfurization which is
normally performed in bioreactors are the most environmentally friendly and eco-
nomically favourable approach, especially compared with their physicochemical
competitors (Syed et al. 2006; Xinqiao et al. 2003; Zăbavă et al. 2019). This can
be attributed to the fact that bio-desulfurization technology normally doesn’t need a
high amount of energy and supplementary chemical compounds (Tabatabaei and
Ghanavati 2018; Zăbavă et al. 2019). Besides, the recovery of sulfur is also possible
so that the recovered sulfur in different forms (depending on the availability of
oxygen in the process) can substitute its chemical-based counterpart (Ramos et al.
2013). This brings more environmental savings and therefore can compensate the
environmental impacts caused by the consumption of energy resources.
Having considered the physicochemical biogas upgrading technologies, the use
of chemicals significantly contributes to the environmental impacts of the upgrading
technology. In other words, the background emissions caused by the production of
chemicals and materials are the most important contributors to the environmental
burdens of physicochemical biogas desulfurization technologies (Zulkefli et al.
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 373
2016). For instance, in chemical scrubbing and immobilized activated carbon desul-
furization, the H2O2 production stage as well as active carbon production and
preparation imposed the highest impacts on the biogas upgrading facility (Cano
et al. 2018).
In order to inject biogas into the natural gas grid or utilize it as transportation fuel,
apart from the H2S, the CO2 content of biogas should also be removed to a great
extent. The existing technologies for upgrading the biogas into the natural gas grade
biomethane, are chiefly divided into physicochemical methods and methanation
technologies (Sun et al. 2015). In the former, the CO2 content of biogas is separated
while in the latter the CO2 components of biogas react with H2 in a Sabatier reaction
either biologically or chemically and thereby forms additional CH4 (Angelidaki et al.
2018). Among the diverse types of physicochemical upgrading technologies, pres-
sure swing adsorption, water scrubbing, amine scrubbing, membrane separation, and
cryogenic separation are the most widely applied technologies which have been
commercialized and long been employed in biogas plants (Kapoor et al. 2019)
The previous LCA studies with focus on biogas upgrading technologies have
shown that biogas upgrading units are among the top two contributors to the
environmental impacts of AD systems (Hauser 2017; Wang et al. 2016). Corre-
spondingly, many studies have been conducted to compare and introduce the most
environmental friendly biogas upgrading technologies (Ardolino and Arena 2019;
Florio et al. 2019; Lombardi and Francini 2020; Starr et al. 2012). Table 14.3
summarizes some of those studies dedicated to the LCA of biogas upgrading
technologies. The results of LCA studies over the biogas upgrading technologies
demonstrated “win-win” situations compared to the fossil routes for the production
of transportation fuel and natural gas. However, comparing different forms of
upgrading technologies have shown that each technology has specific merits and
drawbacks over the others. This attributes to the difference between the input-output
analysis of each technology, including energy, material and emissions.
Methane slippage is the main contributor to the global warming damage category
among different biogas upgrading technologies (Ardolino et al. 2020). Chemical
absorption has the least methane leakage as illustrated in Table 14.4; however, this
facility requires high amounts of heat and further addition of chemicals such as
amines and alkali solutions (Angelidaki et al. 2018). Correspondingly, chemical
absorption would have high background emissions associated with the heat and
chemicals production. Generally, the results of LCA studies of physicochemical
technologies depend on hypothesis made by authors over methane slippage, energy
and chemical consumption. This makes it difficult to directly compare the results of
the previous studies (Ardolino et al. 2020; Hauser 2017; Lombardi and Francini
2020). Furthermore, another parameter which affects the LCA of physicochemical
upgrading technologies, is the possible recovery of CO2 to substitute its conven-
tional industrial counterpart. Although the CO2 released into the environment from
374
a
Value includes required electricity for upgrading, drying and compressing
b
H2S removal is required for high concentrations of sulfur (500 mg H2S/m3 biogas)
c
Recovering CO2 needs energy consumption
d
98% purified
375
(continued)
e
Indirect electricity consumption of 5.94 kWhe for the production of 0.11 kg hydrogen which is required for upgrading 1 kg biogas, and direct electricity
376
upgrading processes is accounted as biogenic, its reuse for industrial purposes brings
more environmental benefits. As illustrated in Table 14.4, recovering highly purified
CO2 from cryogenic separation technology without further energy demand for its
conditioning, could compensate the utilization of high amounts of electricity and
glycerol refrigerant (Hauser 2017).
Biogas methanation technologies referred to as “power to gas” have been intro-
duced as a promising approach to integrate biogas plants in the renewable electricity
sector, which could overcome the challenges of surplus electricity production from
renewable sources (Duan et al. 2020). This novel technology could form a closed-
loop cycle under the concept of circular bioeconomy in which: (1) hydrogen is
converted into biomethane with higher energy content, i.e., the energy content of
biomethane (36 MJ/m3) is almost three times as much as that of hydrogen (10.88 MJ/
m3) (Luo et al. 2012) and (2) the CO2 fraction of biogas is also transformed into an
energy product instead of releasing to the atmosphere without any implication.
However, the greenhouse gas (GHG) emission intensity of the electricity used for
the upstream water electrolysis technology supplying the hydrogen demand of the
methanation process plays a key role in the sustainability of this novel biogas
upgrading method (Zhang et al. 2020). Therefore, if the renewable share of electric-
ity utilized in electrolyzer units is less than 80%, the methanation process at the
downstream leads to higher impacts in all environmental damage categories, com-
pared to physicochemical biogas upgrading processes (Lorenzi et al. 2019).
Apart from the GHG intensity of the electricity, the type of feedstock used for the
production of biomethane at the initial stage is a predominant criterion since biomass
with higher methane potential decreases the hydrogen demand for biogas upgrading
and thereby increases the environmental benefits of the methanation process. More
specifically, in case of chemical methanation a part of the environmental burden is
attributed to the chemical catalyst, alumina supported Nickel-based catalyst
(Ni-α-Al2O3), utilized in the methanation reactor (Lorenzi et al. 2019). It is worth
noting that, if oxygen, which is produced as a co-product in electrolyzer units is
given credit, the net environmental impacts from the AD systems will decrease. This
is due to the fact that the oxygen demand is growing in municipal solid waste (MSW)
and wastewater treatment plants, as well as electric furnace and glass melting
industries and thereby utilizing high amounts of oxygen instead of its prevailing
conventional counterpart from cryogenic air separation that would provide eco-
nomic, environmental and energetic merits (Kato 2007; Khoshnevisan et al. 2020b).
The final use of biogas and digestate and the products they substitute are key
parameters which can affect the overall LCA results. Some of the previous studies
demonstrated that if biogas is used as transportation fuel to substitute fossil-based
378 S. N. Elyasi et al.
Apart from biogas, the digestate remaining at the end of AD process contains
reasonable amounts of macro- and micro-nutrients and is thereby considered as a
substitution for chemical fertilizers (Styles et al. 2018). In fact, utilizing the
remaining slurry after AD as soil amendment not only avoids the generated emis-
sions from the production of chemical fertilizer but also from its application.
However, the limitations such as long storage time, emissions from storage systems,
oversupply of digestate on the farmlands, and the application out of cropping season
or at improper time undermine the sustainability of using digestate as a substitute for
chemical fertilizers (Khoshnevisan et al. 2021). In the previous LCA studies,
ammonia and methane emissions from digestate storage and application as well as
emissions caused by the transportation of such bulky materials imposed the highest
environmental burdens associated with digestate management (Tsapekos et al.
2021). Although transportation of digestate is still challenging, some studies dem-
onstrated that application of digestate on surrounding farmlands would still have
better environmental performance than their use for microalgae or powder
biofertilizer production (Duan et al. 2020)
Depending on the feedstock used for biogas production, the digestate may contain
heavy metals or soil contaminants such as Cu, Zn, and Mn, hence increase the risk of
soil contamination (Valeur 2011). Therefore, digestate may need some pretreatment
to remove contaminants and hazardous components from the digestate before field
application. Moreover, in regions with intensive biogas plants, if long distance
transportation is needed, composting and pelletizing are widely employed as post-
treatment of the digestate (Rehl and Müller 2011). The type of technology used for
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 379
digestate treatment is related to the composition of the remaining slurry from the AD
reactor (Styles et al. 2018). Normally digestate management is carried out through
mechanical, physicochemical or biological processes (Monfet et al. 2018). The
characteristics of each type of technology employed for the post-treatment of
digestate have been thoroughly described in the review paper by Monfet et al.
(2018). From an LCA point of view, the high energy demand of post-treatment
technologies has been found as the environmental hotspot. Mechanical treatment,
i.e., belt drying and thermal concentration have the highest energy demand which
affects the environmental performance of AD systems (Rehl and Müller 2011).
Furthermore, the consumption of specific amounts of chemicals such as sulfuric
acid, sodium hydroxide and powered polymers in physicochemical and biological
treatments would further increase the environmental damages caused by the post-
treatment of the digestate (Vázquez-Rowe et al. 2015).
Fig. 14.3 A simplified process design of an AD based biorefinery. SCP, single cell protein; CHP,
combined heat and power generation; Ag waste, agricultural waste
It is worth noting that the biorefining approach can have a positive influence on
one aspect, but a negative point in another aspect. Holistically, investigating the
LCA and environmental impacts of biorefineries have shown that the cascading use
of biomaterials through integral bioprocesses would lead to economic growth,
positive environmental impacts, expansion of fossil-free bioprocesses, and positive
social effects in terms of improving the regional economy following the product
diversification (Bello et al. 2018; García-Bustamante et al. 2018; Husgafvel et al.
2017). Specifically, the production of various bioproducts through biorefinery path-
ways lead to climate mitigation compared with their chemical processes
(Khoshnevisan et al. 2020b; Sadhukhan and Martinez-Hernandez 2017; Seghetta
et al. 2016). Furthermore, under the concept of a biorefinery, the worthless materials
are returned back into the supply chain more efficiently. To illustrate, by applying
the digestate and biogas downstream for the production of single cell protein (SCP),
the usage of environmental-costly protein sources such as soybean meal would be
limited (Khoshnevisan et al. 2018b).
The exploitation of a wide range of bioproducts under the biorefinery concept
requires higher quantities of energy and additive chemicals/materials (Bello et al.
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 381
Table 14.5 LCA studies focusing on the valorization of various types of biofeedstock to diverse
bioproducts
Type of
biomass Products Remark on LCA Reference
Seaweed Ethanol, proteins and liquid Environmental restoration & Seghetta et al.
fertilizer climate mitigation (2016)
The major environmental
hotspot is energy consumption
Residual Bioethanol, acetic acid, lig- Identified hotspot: Bello et al.
beech nin, and furfural; Glucose, Pre-treatment of biomass; (2018)
woodchip hemicellulose energy demands of the process
and enzyme production
Municipal Levulinic acid, electricity, Global saving of 0.4 kg CO2-eq Sadhukhan
solid waste fertilizer per kg levulinic acid and Martinez-
Hernandez
(2017)
Organic Succinic acid, lactic acid, Biorefining pathway and biogas Khoshnevisan
fraction of single cell protein, usage are the determinable et al. (2020b)
municipal biomethane, heat, electricity, effect on LCA results
solid waste biofertilizer
Vine shoots Lactic acid, and furfural Production of bio-based lactic Pachón et al.
acid or furfural has strong (2020)
environmental benefits
Castor Electricity, biodiesel, etha- Improvement effect on global Khoshnevisan
nol, heat, biomethane, warming and resources; Nega- et al. (2018a)
glycerol tive environmental impacts on
the Human Health and Ecosys-
tem Quality
Citrus peel Phenolic compounds, oil, Replacement of microwave and Joglekar et al.
methane, syngas, ethanol ultrasound-assisted technolo- (2019)
gies by conventional intensify-
ing technology improve the
environmental effects
Sugars and Succinic acid, adipic acid, The life-cycle GHG results are Cai et al.
lignin Biodiesel highly sensitive to co-product (2018)
handling methods
Wheat straw Bioethanol, lactic acid, elec- Net environmental saving com- Parajuli et al.
& alfalfa tricity, feed protein, pared to petrochemical compet- (2017)
biofertilizer, fodder silage itors of bioproducts
Energy consumption, specifi-
cally for the production of
enzymes in biorefinery was a
major hotspot
Olive wastes Biofuel, phosphate salts, Mitigation of environmental Khounani
natural antioxidant, and an impacts of foods et al. (2021)
oxygenated fuel additive Transition from olive agri-food
(triacetin) to agro-biorefinery is environ-
mentally favorable
382 S. N. Elyasi et al.
2018). For instance, the production of lactic acid under a biorefinery approach
requires high energy and chemical consumption (Liu et al. 2021). Thus, the choice
of biorefinery technologies in terms of energy and pollutant intensity have major
environmental effects (Bello et al. 2018; Seghetta et al. 2016). Such limitations are
the reason that the production of biofuel still outperforms the production of bio-
chemicals (Khoshnevisan et al. 2018a, 2020b). This is due to the low conversion
efficiency as well as the energy intensity of biorefining technologies (Bello et al.
2018; Khoshnevisan et al. 2020b; Seghetta et al. 2016). Overall, downstream biogas
usage, bioproduct substitutions, and the conversion rate are the most effective
parameters affecting the LCA of biorefineries (Khoshnevisan et al. 2020b). If
agricultural biomass is used in biorefineries, some other criteria such as crop
cultivation and land use change should not be neglected in the boundary of LCA
studies (Katakojwala and Mohan 2020; Liu et al. 2021)
Despite all the attempts made so far to investigate the LCA of biorefineries, it still
suffers from low transparency and clarity. Hence, future research should critically
evaluate the environmental impacts of the AD process under the biorefinery concept
to attain more sophisticated knowledge and thereby selecting the most suitable
interlocked bioprocesses for biogas production. Also, further attempts should be
made for process optimization to satisfy the environmental indicators. This would
lead to the production of more value-added products alongside bioenergy and thus
overcoming the environmental hazards of the petrochemical based production of
building block chemicals and materials.
References
Adelt M, Wolf D, Vogel A (2011) LCA of biomethane. J Nat Gas Sci Eng 3:646–650
Adnan AI, Ong MY, Nomanbhay S, Chew KW, Show PL (2019) Technologies for biogas
upgrading to biomethane: a review. Bioengineering 6:92. https://doi.org/10.3390/
bioengineering6040092
Angelidaki I, Treu L, Tsapekos P, Luo G, Campanaro S, Wenzel H, Kougias PG (2018) Biogas
upgrading and utilization: current status and perspectives. Biotechnol Adv 36:452–466
Ardolino F, Arena U (2019) Biowaste-to-biomethane: an LCA study on biogas and syngas roads.
Waste Manag 87:441–453
Ardolino F, Parrillo F, Arena U (2018) Biowaste-to-biomethane or biowaste-to-energy? An LCA
study on anaerobic digestion of organic waste. J Clean Prod 174:462–476
Ardolino F, Cardamone GF, Parrillo F, Arena U (2020) Biogas-to-biomethane upgrading: a
comparative review and assessment in a life cycle perspective. Renew Sust Energ Rev 110588
Asgher M, Bashir F, Iqbal HMN (2014) A comprehensive ligninolytic pre-treatment approach from
lignocellulose green biotechnology to produce bio-ethanol. Chem Eng Res Des 92:1571–1578
Aziz NIHA, Hanafiah MM, Gheewala SH (2019) A review on life cycle assessment of biogas
production: challenges and future perspectives in Malaysia. Biomass Bioenergy 122:361–374
Bacenetti J, Sala C, Fusi A, Fiala M (2016) Agricultural anaerobic digestion plants: what LCA
studies pointed out and what can be done to make them more environmentally sustainable. Appl
Energy 179:669–686
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 383
Di Maria F, Micale C, Contini S (2016) A novel approach for uncertainty propagation applied to
two different bio-waste management options. Int J Life Cycle Assess 21:1529–1537
Duan N, Khoshnevisan B, Lin C, Liu Z, Liu H (2020) Life cycle assessment of anaerobic digestion
of pig manure coupled with different digestate treatment technologies. Environ Int 137:105522
Edwards J, Othman M, Crossin E, Burn S (2017) Anaerobic co-digestion of municipal food waste
and sewage sludge: a comparative life cycle assessment in the context of a waste service
provision. Bioresour Technol 223:237–249
Elyasi SN, He L, Tsapekos P, Rafiee S, Khoshnevisan B, Carbajales-Dale M, Mohtasebi SS, Liu H,
Angelidaki I (2021) Could biological biogas upgrading be a sustainable substitution for water
scrubbing technology? A case study in Denmark. Energy Convers Manag 245:114550. https://
doi.org/10.1016/j.enconman.2021.114550
Eriksson O, Bisaillon M, Haraldsson M, Sundberg J (2016) Enhancement of biogas production
from food waste and sewage sludge—environmental and economic life cycle performance. J
Environ Manag 175:33–39
Ertem FC, Martínez-Blanco J, Finkbeiner M, Neubauer P, Junne S (2016) Life cycle assessment of
flexibly fed biogas processes for an improved demand-oriented biogas supply. Bioresour
Technol 219:536–544. https://doi.org/10.1016/j.biortech.2016.07.123
Finkbeiner M (2014) The international standards as the constitution of life cycle assessment: the
ISO 14040 series and its offspring. In: Background and future prospects in life cycle assessment.
Springer, pp 85–106
Florio C, Fiorentino G, Corcelli F, Ulgiati S, Dumontet S, Güsewell J, Eltrop L (2019) A life cycle
assessment of biomethane production from waste feedstock through different upgrading tech-
nologies. Energies 12
Fusi A, Bacenetti J, Fiala M, Azapagic A (2016) Life cycle environmental impacts of electricity
from biogas produced by anaerobic digestion. Front Bioeng Biotechnol 4:26
García-Bustamante CA, Aguilar-Rivera N, Zepeda-Pirrón M, Armendáriz-Arnez C (2018) Devel-
opment of indicators for the sustainability of the sugar industry. Environ Socio-Econ Stud 6:22–
38
Gnaoui YE, Karouach F, Bakraoui M, Barz M, Bari HE (2020) Mesophilic anaerobic digestion of
food waste: effect of thermal pretreatment on improvement of anaerobic digestion process.
Energy Rep 6:417–422
Hauser MJ (2017) Cost evaluation and life cycle assessment of biogas upgrading technologies for
an anaerobic digestion case study in the United States. NTNU
Hernández-Beltrán JU, Lira H-D, Omar I, Cruz-Santos MM, Saucedo-Luevanos A, Hernández-
Terán F, Balagurusamy N (2019) Insight into pretreatment methods of lignocellulosic biomass
to increase biogas yield: current state, challenges, and opportunities. Appl Sci 9:3721
Hetemäki L, Hanewinkel M, Muys B, Ollikainen M, Palahí M, Trasobares A, Aho E, Ruiz CN,
Persson G, Potoćnik J (2017) Leading the way to a European circular bioeconomy strategy.
European Forest Institute
Hijazi O, Munro S, Zerhusen B, Effenberger M (2016) Review of life cycle assessment for biogas
production in Europe. Renew Sustain Energy Rev 54:1291–1300
Hoyer K, Hulteberg C, Svensson M, Jernberg J, Nörregård Ö (2016) Biogas upgrading-technical
review. Energiforsk, Stockholm
Husgafvel R, Poikela K, Honkatukia J, Dahl O (2017) Development and piloting of sustainability
assessment metrics for arctic process industry in Finland—the biorefinery investment and slag
processing service cases. Sustainability 9:1693
Jacobson MZ, Delucchi MA, Bazouin G, Bauer ZA, Heavey CC, Fisher E, Morris SB, Piekutowski
DJ, Vencill TA, Yeskoo TW (2015) 100% clean and renewable wind, water, and sunlight
(WWS) all-sector energy roadmaps for the 50 United States. Energy Environ Sci 8:2093–2117
Jayachandra T, Venugopal C, Appaiah KA (2011) Utilization of phytotoxic agro waste—Coffee
cherry husk through pretreatment by the ascomycetes fungi Mycotypha for biomethanation.
Energy Sustain Dev 15:104–108
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 385
Jensen MB, Møller J, Scheutz C (2016) Comparison of the organic waste management systems in
the Danish–German border region using life cycle assessment (LCA). Waste Manag 49:491–
504
Joglekar SN, Pathak PD, Mandavgane SA, Kulkarni BD (2019) Process of fruit peel waste
biorefinery: a case study of citrus waste biorefinery, its environmental impacts and recommen-
dations. Environ Sci Pollut Res 26:34713–34722
Kampman B, Leguijt C, Scholten T, Tallat-Kelpsaite J, Brückmann R, Maroulis G, Lesschen JP,
Meesters K, Sikirica N, Elbersen B, (2017) Optimal use of biogas from waste streams: an
assessment of the potential of biogas from digestion in the EU beyond 2020
Kapoor R, Ghosh P, Kumar M, Vijay VK (2019) Evaluation of biogas upgrading technologies and
future perspectives: a review. Environ Sci Pollut Res 26:11631–11661
Karuppiah T, Azariah VE (2019) Biomass pretreatment for enhancement of biogas production. In:
Anaerobic digestion. IntechOpen, London
Katakojwala R, Mohan SV (2020) A critical view on the environmental sustainability of biorefinery
systems. Curr Opin Green Sustain Chem 100392
Kato T (2007) Effective utilization of by-product oxygen of electrolysis hydrogen production.
Nagoya
Khoshnevisan B, Shafiei M, Rajaeifar MA, Tabatabaei M (2016) Biogas and bioethanol production
from pinewood pre-treated with steam explosion and N-methylmorpholine-N-oxide (NMMO):
a comparative life cycle assessment approach. Energy 114:935–950
Khoshnevisan B, Rafiee S, Tabatabaei M, Ghanavati H, Mohtasebi SS, Rahimi V, Shafiei M,
Angelidaki I, Karimi K (2018a) Life cycle assessment of castor-based biorefinery: a well to
wheel LCA. Int J Life Cycle Assess 23:1788–1805
Khoshnevisan B, Tsapekos P, Alvarado-Morales M, Rafiee S, Tabatabaei M, Angelidaki I (2018b)
Life cycle assessment of different strategies for energy and nutrient recovery from source sorted
organic fraction of household waste. J Clean Prod 180:360–374
Khoshnevisan B, Dodds M, Tsapekos P, Torresi E, Smets BF, Angelidaki I, Zhang Y, Valverde-
Pérez B (2020a) Coupling electrochemical ammonia extraction and cultivation of methane
oxidizing bacteria for production of microbial protein. J Environ Manag 265:110560
Khoshnevisan B, Tabatabaei M, Tsapekos P, Rafiee S, Aghbashlo M, Lindeneg S, Angelidaki I
(2020b) Environmental life cycle assessment of different biorefinery platforms valorizing
municipal solid waste to bioenergy, microbial protein, lactic and succinic acid. Renew Sust
Energ Rev 117:109493
Khoshnevisan B, Duan N, Tsapekos P, Awasthi MK, Liu Z, Mohammadi A, Angelidaki I, Tsang
DCW, Zhang Z, Pan J, Ma L, Aghbashlo M, Tabatabaei M, Liu H (2021) A critical review on
livestock manure biorefinery technologies: sustainability, challenges, and future perspectives.
Renew Sust Energ Rev 135:110033
Khounani Z, Hosseinzadeh-Bandbafha H, Moustakas K, Talebi AF, Goli SAH, Rajaeifar MA,
Khoshnevisan B, Salehi Jouzani G, Peng W, Kim K-H, Aghbashlo M, Tabatabaei M, Lam SS
(2021) Environmental life cycle assessment of different biorefinery platforms valorizing olive
wastes to biofuel, phosphate salts, natural antioxidant, and an oxygenated fuel additive
(triacetin). J Clean Prod 278:123916
Kohlheb N, Wluka M, Bezama A, Thrän D, Aurich A, Müller RA (2020) Environmental-economic
assessment of the pressure swing adsorption biogas upgrading technology. BioEnergy Res:1–9
Koupaie EH, Dahadha S, Lakeh AB, Azizi A, Elbeshbishy E (2019) Enzymatic pretreatment of
lignocellulosic biomass for enhanced biomethane production-A review. J Environ Manag 233:
774–784
Kral I, Piringer G, Saylor MK, Gronauer A, Bauer A (2016) Environmental effects of steam
explosion pretreatment on biogas from maize—case study of a 500-kW Austrian biogas facility.
BioEnergy Res 9:198–207
Lam C-M, Iris K, Hsu S-C, Tsang DC (2018) Life-cycle assessment on food waste valorisation to
value-added products. J Clean Prod 199:840–848
386 S. N. Elyasi et al.
Lee JTE, Khan MU, Tian H, Ee AW, Lim EY, Dai Y, Tong YW, Ahring BK (2020) Improving
methane yield of oil palm empty fruit bunches by wet oxidation pretreatment: mesophilic and
thermophilic anaerobic digestion conditions and the associated global warming potential effects.
Energy Convers Manag 225:113438
Leonzio G (2016) Upgrading of biogas to bio-methane with chemical absorption process: simula-
tion and environmental impact. J Clean Prod 131:364–375
Lima PDM, Colvero DA, Gomes AP, Wenzel H, Schalch V, Cimpan C (2018) Environmental
assessment of existing and alternative options for management of municipal solid waste in
Brazil. Waste Manag 78:857–870
Liu Y, Lyu Y, Tian J, Zhao J, Ye N, Zhang Y, Chen L (2021) Review of waste biorefinery
development towards a circular economy: from the perspective of a life cycle assessment.
Renew Sust Energ Rev 139:110716
Lombardi L, Francini G (2020) Techno-economic and environmental assessment of the main biogas
upgrading technologies. Renew Energy 156:440–458
Lorenzi G, Gorgoroni M, Silva C, Santarelli M (2019) Life Cycle Assessment of biogas upgrading
routes. Energy Procedia 158:2012–2018
Luo G, Johansson S, Boe K, Xie L, Zhou Q, Angelidaki I (2012) Simultaneous hydrogen utilization
and in situ biogas upgrading in an anaerobic reactor. Biotechnol Bioeng 109:1088–1094
Martínez-Gutiérrez E (2018) Biogas production from different lignocellulosic biomass sources:
advances and perspectives. 3 Biotech 8:1–18
McCarty PL (1982) One-hundred years of anaerobic treatment. In: Hughes DE, Stafford DA (eds)
Anaerobic digestion. Elsevier Biomedical Press, New York
Mills N, Pearce P, Farrow J, Thorpe R, Kirkby N (2014) Environmental & economic life cycle
assessment of current & future sewage sludge to energy technologies. Waste Manag 34:185–
195
Moioli S, Hijazi O, Pellegrini LA, Bernhardt H (2020) Simulation of different biogas upgrading
processes and LCA for the selection of the best technology. In: 2020 ASABE annual interna-
tional virtual meeting. American Society of Agricultural and Biological Engineers, p 1
Monfet E, Aubry G, Ramirez AA (2018) Nutrient removal and recovery from digestate: a review of
the technology. Biofuels 9:247–262
Montgomery LF, Bochmann G (2014) Pretreatment of feedstock for enhanced biogas production.
IEA Bioenergy Ireland
Morero B, Groppelli E, Campanella EA (2015) Life cycle assessment of biomethane use in
Argentina. Bioresour Technol 182:208–216
Mudhoo A (2012) Biogas production: pretreatment methods in anaerobic digestion. Wiley,
Hoboken, NJ
Nayal FS, Mammadov A, Ciliz N (2016) Environmental assessment of energy generation from
agricultural and farm waste through anaerobic digestion. J Environ Manag 184:389–399
Nizami AS, Rehan M, Waqas M, Naqvi M, Ouda OKM, Shahzad K, Miandad R, Khan MZ,
Syamsiro M, Ismail IMI, Pant D (2017) Waste biorefineries: enabling circular economies in
developing countries. Bioresour Technol 241:1101–1117
Pachón ER, Mandade P, Gnansounou E (2020) Conversion of vine shoots into bioethanol and
chemicals: prospective LCA of biorefinery concept. Bioresour Technol 303:122946
Parajuli R, Knudsen MT, Birkved M, Djomo SN, Corona A, Dalgaard T (2017) Environmental
impacts of producing bioethanol and biobased lactic acid from standalone and integrated
biorefineries using a consequential and an attributional life cycle assessment approach. Sci
Total Environ 598:497–512
Pérez-Camacho MN, Curry R, Cromie T (2019) Life cycle environmental impacts of biogas
production and utilisation substituting for grid electricity, natural gas grid and transport fuels.
Waste Manag 95:90–101
Pericault Y, Risberg M, Vesterlund M, Viklander M, Hedström A (2017) A novel freeze protection
strategy for shallow buried sewer pipes: temperature modelling and field investigation. Water
Sci Technol 76:294–301
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 387
Piao W, Kim Y, Kim H, Kim M, Kim C (2016) Life cycle assessment and economic efficiency
analysis of integrated management of wastewater treatment plants. J Clean Prod 113:325–337
Pilli S, Pandey AK, Katiyar A, Pandey K, Tyagi RD (2020) Pre-treatment technologies to enhance
anaerobic digestion, sustainable sewage sludge management and resource efficiency.
IntechOpen
Prasad A, Sotenko M, Blenkinsopp T, Coles SR (2016) Life cycle assessment of lignocellulosic
biomass pretreatment methods in biofuel production. Int J Life Cycle Assess 21:44–50
Rajendran K, Murthy GS (2019) Techno-economic and life cycle assessments of anaerobic
digestion—a review. Biocatal Agric Biotechnol 20:101207
Rajput AA, Zeshan, Visvanathan C (2018) Effect of thermal pretreatment on chemical composition,
physical structure and biogas production kinetics of wheat straw. J Environ Manag 221:45–52
Ramos I, Pérez R, Fdz-Polanco M (2013) Microaerobic desulphurisation unit: a new biological
system for the removal of H2S from biogas. Bioresour Technol 142:633–640
Rao PV, Baral SS, Dey R, Mutnuri S (2010) Biogas generation potential by anaerobic digestion for
sustainable energy development in India. Renew Sust Energ Rev 14:2086–2094
Ravindran R, Jaiswal S, Abu-Ghannam N, Jaiswal AK (2018) A comparative analysis of
pretreatment strategies on the properties and hydrolysis of brewers’ spent grain. Bioresour
Technol 248:272–279
Rehl T, Müller J (2011) Life cycle assessment of biogas digestate processing technologies. Resour
Conserv Recycl 56:92–104
Rehl T, Müller J (2013) CO2 abatement costs of greenhouse gas (GHG) mitigation by different
biogas conversion pathways. J Environ Manag 114:13–25
Righi S, Bandini V, Marazza D, Baioli F, Torri C, Contin A (2016) Life cycle assessment of high
ligno-cellulosic biomass pyrolysis coupled with anaerobic digestion. Bioresour Technol 212:
245–253
Sadhukhan J, Martinez-Hernandez E (2017) Material flow and sustainability analyses of biorefining
of municipal solid waste. Bioresour Technol 243:135–146
Santos-Clotas E, Cabrera-Codony A, Castillo A, Martín MJ, Poch M, Monclús H (2019) Environ-
mental decision support system for biogas upgrading to feasible fuel. Energies 12:1546
Seghetta M, Hou X, Bastianoni S, Bjerre A-B, Thomsen M (2016) Life cycle assessment of
macroalgal biorefinery for the production of ethanol, proteins and fertilizers—a step towards a
regenerative bioeconomy. J Clean Prod 137:1158–1169
Shimako AH, Tiruta-Barna L, Pigné Y, Benetto E, Gutiérrez TN, Guiraud P, Ahmadi A (2016)
Environmental assessment of bioenergy production from microalgae based systems. J Clean
Prod 139:51–60
Shrestha B, Hernandez R, Fortela DLB, Sharp W, Chistoserdov A, Gang D, Revellame E,
Holmes W, Zappi ME (2020) A Review of pretreatment methods to enhance solids reduction
during anaerobic digestion of municipal wastewater sludges and the resulting digester perfor-
mance: implications to future urban biorefineries. Appl Sci 10:9141
Smullen E, Finnan J, Dowling D, Mulcahy P (2017) Bioconversion of switchgrass: identification of
a leading pretreatment option based on yield, cost and environmental impact. Renew Energy
111:638–645
Smullen E, Finnan J, Dowling D, Mulcahy P (2019) The environmental performance of
pretreatment technologies for the bioconversion of lignocellulosic biomass to ethanol. Renew
Energy 142:527–534
Song Z, Yang G, Guo Y, Zhang T (2012) Comparison of two chemical pretreatments of rice straw
for biogas production by anaerobic digestion. Bioresources 7:3223–3236
Starr K, Gabarrell X, Villalba G, Talens L, Lombardi L (2012) Life cycle assessment of biogas
upgrading technologies. Waste Manag 32:991–999
Strazza C, Magrassi F, Gallo M, Del Borghi A (2015) Life Cycle Assessment from food to food: a
case study of circular economy from cruise ships to aquaculture. Sustain Prod Consumpt 2:40–
51
388 S. N. Elyasi et al.
Styles D, Adams P, Thelin G, Vaneeckhaute Cl, Chadwick D, Withers PJ (2018) Life cycle
assessment of biofertilizer production and use compared with conventional liquid digestate
management. Environ Sci Technol 52:7468–7476
Sun Q, Li H, Yan J, Liu L, Yu Z, Yu X (2015) Selection of appropriate biogas upgrading
technology-a review of biogas cleaning, upgrading and utilisation. Renew Sust Energ Rev 51:
521–532
Syed M, Soreanu G, Falletta P, Béland M (2006) Removal of hydrogen sulfide from gas streams
using biological processes—a review. Can Biosyst Eng 48:2
Tabatabaei M, Ghanavati H (2018) Biogas: fundamentals, process, and operation. Springer, Cham
Tabatabaei M, Aghbashlo M, Valijanian E, Panahi HKS, Nizami A-S, Ghanavati H, Sulaiman A,
Mirmohamadsadeghi S, Karimi K (2020) A comprehensive review on recent biological inno-
vations to improve biogas production, part 1: upstream strategies. Renew Energy 146:1204–
1220
Torres ML, Lloréns MCE (2008) Effect of alkaline pretreatment on anaerobic digestion of solid
wastes. Waste Manag 28:2229–2234
Tsapekos P, Khoshnevisan B, Alvarado-Morales M, Zhu X, Pan J, Tian H, Angelidaki I (2021)
Upcycling the anaerobic digestion streams in a bioeconomy approach: A review. Renew Sustain
Energy Rev 15:1111635. https://doi.org/10.1016/j.rser.2021.111635
Valeur I (2011) Speciation of heavy metals and nutrient elements in digestate. Norwegian Univer-
sity of Life Sciences, Ås
Vannarath A, Thalla AK (2019) Evaluation, ranking, and selection of pretreatment methods for the
conversion of biomass to biogas using multi-criteria decision-making approach. Environ Syst
Decis 1–16
Vázquez-Rowe I, Golkowska K, Lebuf V, Vaneeckhaute C, Michels E, Meers E, Benetto E, Koster
D (2015) Environmental assessment of digestate treatment technologies using LCA methodol-
ogy. Waste Manag 43:442–459
Wang D, Ai P, Yu L, Tan Z, Zhang Y (2015) Comparing the hydrolysis and biogas production
performance of alkali and acid pretreatments of rice straw using two-stage anaerobic fermenta-
tion. Biosyst Eng 132:47–55
Wang Q-L, Li W, Gao X, Li S-J (2016) Life cycle assessment on biogas production from straw and
its sensitivity analysis. Bioresour Technol 201:208–214
Wang S, Yu S, Lu Q, Liao Y, Li H, Sun L, Wang H, Zhang Y (2020) Development of an alkaline/
acid pre-treatment and anaerobic digestion (APAD) process for methane generation from waste
activated sludge. Sci Total Environ 708:134564
Xiao C, Ma Y, Ji D, Zang L (2017) Review of desulfurization process for biogas purification. In:
IOP conference series: earth and environmental science. IOP, p 012177
Xiao C, Fu Q, Liao Q, Huang Y, Xia A, Chen H, Zhu X (2020) Life cycle and economic
assessments of biogas production from microalgae biomass with hydrothermal pretreatment
via anaerobic digestion. Renew Energy 151:70–78
Xinqiao Z, Bing G, Jinshen D (2003) Study on removal of hydrogen sulfide from industrial acidic
waste gas by immobilized microorganism. Spec Petrochem 02
Zăbavă B-Ș, Voicu G, Ungureanu N, Dincă M, Paraschiv G, Munteanu M, Ferdes M (2019)
Methods of biogas purification—a review
Zha X, Tsapekos P, Zhu X, Khoshnevisan B, Lu X, Angelidaki I (2021) Bioconversion of
wastewater to single cell protein by methanotrophic bacteria. Bioresour Technol 320:124351
14 Up and Downstream Technologies of Anaerobic Digestion from Life Cycle. . . 389
Zhang Q, Tang L, Zhang J, Mao Z, Jiang L (2011) Optimization of thermal-dilute sulfuric acid
pretreatment for enhancement of methane production from cassava residues. Bioresour Technol
102:3958–3965
Zhang X, Witte J, Schildhauer T, Bauer C (2020) Life cycle assessment of power-to-gas with biogas
as the carbon source. Sustain Energy Fuels 4:1427–1436
Zhen G, Lu X, Kato H, Zhao Y, Li Y-Y (2017) Overview of pretreatment strategies for enhancing
sewage sludge disintegration and subsequent anaerobic digestion: current advances, full-scale
application and future perspectives. Renew Sust Energ Rev 69:559–577
Zulkefli N, Masdar M, Jahim J, Harianto E (2016) Overview of H2S removal technologies from
biogas production. Int J Appl Eng Res 11:10060–10100
Chapter 15
Life Cycle Assessment of Anaerobic
Digestion Systems: An Approach Towards
Sustainable Waste Management
Abstract The wastes generated from different human activities such as agricultural
or crop wastes, livestock, food wastes are biodegradable wastes, which can be
utilized for various purposes including biogas generation and production of other
value added products. The main aim of this chapter is to provide information on the
types of waste that can be used to produce biogas through anaerobic digestion
(AD) system. Anaerobic digestion is a process of conversion of biodegradable
waste into biogas by the action of microbial communities. The gases obtained during
AD mainly consist of methane, carbon-dioxide and some amount of water vapour.
Better understanding of the different processes, including the pretreatment and post-
treatment involved in the AD, is necessary to develop the green technology to
achieve sustainable development. In this regard, LCA is a great tool to study the
feasibility of the AD process for environmental sustainability. In this chapter, an
introduction to life cycle assessment (LCA) and description of previous studies on
the LCA of anaerobic digestion systems towards environmentally sustainable man-
agement practices are presented. The chapter also discusses LCA for improvement
of the performance of the anaerobic digestion process of various biodegradable
wastes for generation of biogas. The findings suggest that more future research is
needed for process improvement and to promote the wide application of the AD
process for generation of green energy.
M. M. Hanafiah (*)
Department of Earth Sciences and Environment, Faculty of Science and Technology, Universiti
Kebangsaan Malaysia, Bangi, Selangor, Malaysia
Centre for Tropical Climate Change System, Institute of Climate Change, Universiti
Kebangsaan Malaysia, Bangi, Selangor, Malaysia
e-mail: [email protected]
I. Ansari · K. Chelvam
Department of Earth Sciences and Environment, Faculty of Science and Technology, Universiti
Kebangsaan Malaysia, Bangi, Selangor, Malaysia
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 391
A. Sinharoy, P. N. L. Lens (eds.), Renewable Energy Technologies for Energy
Efficient Sustainable Development, Applied Environmental Science and Engineering
for a Sustainable Future, https://doi.org/10.1007/978-3-030-87633-3_15
392 M. M. Hanafiah et al.
15.1 Introduction
Biodegradable waste like agricultural and crop residue, food waste, livestock wastes,
sewage sludge, organic municipal waste are rich sources for energy production,
including biogas or biofuels (Appels et al. 2011; Banch et al. 2020b). Conversion of
organic wastes to energy and by-products has become increasingly important for
achieving the circular economy (Banch et al. 2020a). Utilization of such wastes for
bioenergy production provides many advantages such as low-cost of substrate,
minimizing environmental impacts by reducing waste generation, providing extra
revenue to the producers, reducing fossil fuel dependency and making energy
sufficient (Hanafiah et al. 2018; Nizam et al. 2020). Anaerobic digestion systems,
embracing the concept of waste-to-energy is one of the environmentally friendly
technologies for generating biofuel and biogas (Aziz et al. 2019; Vambol et al.
2019).
However, in many areas, solid wastes are being dumped directly to the environ-
ment without proper treatment and management (Ashraf et al. 2019). Issue related to
improper waste management affects human health and the environment (Al-Raad
et al. 2020; Ashraf and Hanafiah 2019; Hanafiah et al. 2020). A common disposal
method such as landfill consumes large amounts of land resources, energy, and
causes significant atmospheric greenhouse gas emissions (Banch et al. 2019). Apart
from that, landfills also produce a substantial amount of landfill leachate containing
high concentrations of organic and inorganic pollutants that poses hazards to the
environment. Therefore, in recent years, research in alternative renewable energy
generated from organic wastes is gaining more interest.
The waste-to-energy concept has been introduced to solve environmental degra-
dation related to waste disposal (Ashraf and Hanafiah 2017). Various technologies
and methods have been used to convert waste into bioenergy. However, this further
requires investigation on the impacts of these technologies on the environment,
human health and economical aspects. In this context, holistic and comprehensive
assessment of the impacts of anaerobic digestion (AD) systems used to convert
waste to energy is important. This will help to identify strategic and feasible
improvements in waste management and renewable energy generation towards
environmental sustainability (Aziz et al. 2020). This chapter provides useful infor-
mation to researchers, students, and policy makers on life cycle assessment (LCA) of
AD systems.
THE BIOCHEMISTRY
Heat • Fertiliser
• Soil amendments Acetic acid 3. Acetogenesis H2 CO2
Biomethane • Livestock bedding
Fuel 4. Methanogenesis
Gas grid CH4 + CO2
Fig. 15.1 Overview of the anaerobic digestion process along with the applications of its products in different sectors
M. M. Hanafiah et al.
15 Life Cycle Assessment of Anaerobic Digestion Systems: An Approach Towards. . . 395
The biogas produced is mostly methane (CH4) and carbon dioxide (CO2) with a
minimal amount of water vapour and other gases. The carbon dioxide and other
gases can be removed leaving only methane which is the primary component of
interest. This biogas purification process will increase the energy value of the biogas
as a source of renewable energy to produce heat and electricity for power engines,
heat generation in digesters, boilers and heating furnaces, run alternative-fuel vehi-
cles and domestic and industrial usage as alternative to natural gases.
Digestate is the other product which is a wet mixture of the material that is left
after the anaerobic digestion. It is separated into solid and liquid which are rich in
nutrients and can be used as fertilisers for crops. Digestate can be applied directly on
the land and added into the soil to improve soil quality which will promote plant
growth. Moreover, some engineering techniques can be applied post-digestion to
recover the phosphorus and nitrogen present in the digestate and produce concen-
trated nutrient products such as struvite (magnesium-ammonium-phosphate) and
ammonium sulfate fertilisers. Other than that, digestate can also be made as bedding
for livestock and be used for soil amendments. Figure 15.2 illustrates the elements of
the biogas recovery system.
Waste generation in recent times is quite high in amount and particularly for food
waste it is up to 32% from all food produced throughout the world (Morales-Polo
et al. 2018). This leads to a remarkable social and economic loss as well as
environmental issues (Browne and Murphy 2013; FAO 2011; Papargyropoulou
et al. 2014). According to the Ministry of Housing and Local Government
(MHLG) Malaysia, disposal of food waste directly to the landfill sites is one of the
main sources of greenhouse gases (GHG) emissions in Malaysia (Hoo et al. 2017).
Another research carried out by Paritosh et al. (2017) also states that the carbon
footprint of food waste is estimated to contribute to the GHG emissions by accu-
mulating approximately 3.3 billion tonnes of CO2 in the atmosphere annually.
Food waste makes up a large percentage of a landfill, where only a very minimal
amount of food waste is being recycled into soil improver or fertiliser. The remaining
food waste in landfills releases methane to the atmosphere as it breaks down.
However, the energy potential produced from food waste is very significant. During
396 M. M. Hanafiah et al.
Biogas
Biogas Handling System
Electricity,Fuel & Flare
Collects and treats biogas excess
recent years, due to the generation of tremendous amounts of food waste, it can be a
promising substrate for the AD process (Lin et al. 2013; Paritosh et al. 2017). Food
waste (FW) poses the characteristics of high moisture and organic contents due to
which it is easily degradable (Li et al. 2019a; Ren et al. 2018). During the AD
process, organic matter present in FW is converted into CH4, CO2 and other gases.
However, FW is characterized by a low C/N ratio (Akindele and Sartaj 2017).
Hence, many times the AD process of FW results in the accumulation of volatile
fatty acids (VFAs) in the reactor, due to the imbalanced nutritional level in the
substrates (Chukwudi et al. 2019; Yazdanpanah et al. 2018).
There are many studies on biogas generation using food waste (Table 15.1).
El-Mashad and Zhang (2010) investigated the production of biogas from food waste
in batch digesters under mesophilic conditions (35 C). The methane produced from
the food waste was 353 L/kg VS after 30 days of digestion. Al-Wahaibi et al. (2020)
carried out a techno-economic investigation of biogas production from different
types of food waste. The methane production during the fermentation period showed
a good correlation between the theoretical and experimental values with a coefficient
of determination R2 ¼ 0.99. The cumulative gas production from mixed food waste
samples was 1550 mL/1 g of dry matter at day 21. Jeong-Ik et al. (2018) have studied
the biogas production from food waste with wood chips via anaerobic co-digestion.
Use of wood chips as co-substrate increased the rate of methane and hydrogen
production in AD. 20 mL/g of methane and 13.9 mL/g of hydrogen were generated
at the food waste to wood chip ratio of 0.5, during 15 days at 35 C.
15 Life Cycle Assessment of Anaerobic Digestion Systems: An Approach Towards. . . 397
Table 15.1 Biogas production from different types of waste using different AD systems
Type of Quantity/
Waste/location Study Method biogas productivity Reference
Fruit and vege- Lab Anaerobic diges- Methane 50–60% in Maile et al.
table waste tion (biochemical volume (2016)
(FVW), Johan- methane potential),
nesburg mar- bioprocess
ket, AMPTS II
South Africa
Kitchen waste Both The batch diges- Methane Productivity Wang et al.
(KW) and fruit/ lab & tion in lab-scale in (0.725L CH4/g (2014)
vegetable waste pilot- 2-L conical flasks, VS)
(FVW), China scale Single-phase
co-digestion
A10-L completely
stirred tank reactor
(CSTR)
Fruit and vege- Lab Single-stage anaer- Methane Methane yield of Farhat et al.
table waste obic sequencing 340 L/kg volatile (2018)
(FVW), Tunisia batch reactors solids (VS)
(ASBRs) and con- Biogas (8% H2,
tinuously stirred 28.5% CO2 and
tank reactor 63.5% CH4)
(CSTR)
Municipal solid Lab Anaerobic Biogas Biogas 493.8 N Pavi et al.
waste (MSW) co-digestion and mL/g VS, meth- (2017)
and fruit and (batch) methane ane- 396.6 N
vegetable waste mL/g VS,
(FVW), Brazil
Fruit and vege- Lab Anaerobic diges- Biogas Biogas: Masebinu
table waste tion semi- and 0.87 Nm3/kg VS et al.
(FVW) Johan- continuous digester methane with 57.58% (2018)
nesburg Mar- methane
ket,
South Africa
Market waste Lab Anaerobic diges- Biogas 35 L/kg MW/day Ranade
(vegetables & tion 200L Biogas et al.
fruits), Pune, plant (1987)
India
Fruit and vege- Lab Anaerobic Methane MSS: 0.303 m3/ Arhoun
table waste co-digestion (kg VS) et al.
(FVW), munici- FVW: 0.403 m3/ (2019)
pal sewage (kg VS), average
sludge (MSS), methane content
Malaga, Spain of digester biogas
was about 62–
64%
Fish waste (FW) Lab Anaerobic Biogas Average biogas Bouallagui
Fruit and vege- co-digestion anaer- production rate et al.
table waste obic sequencing varied between (2009)
(FVW), Tunis, batch reactors 1.53 and 2.53 L/
Tunisia (ASBR) day
(continued)
398 M. M. Hanafiah et al.
Gaby et al. (2017) studied the effect of the retention time and temperature on
microbial community composition and methane production in staged anaerobic
digesters fed with food waste. During this experiment, two anaerobic digestion
parameters were studied, namely the effect of 55 and 65 C as acidogenic reactor
temperature, and the effect of lowering the hydraulic retention time (HRT) from
17 to 10 days in the methanogenic reactor. Higher acetate and butyrate concentra-
tions were found in the 65 C acidogenic reactor. The CH4 production was increased
from ~3600 mL/day to ~7800 mL/day when the HRT was decreased to 10 days. The
effect of yeast on the performance of biogas production from food waste has been
studied (Gao et al. 2020). The result showed that biogas production increased by
520 and 550 mL with addition of 2.0% (of volatile solids; VS) of activated yeast on,
respectively, the 12th and 37th day of anaerobic digestion and the gas production
was relatively stable (Gao et al. 2020).
Pallan et al. (2018) reported that anaerobic digestion of organic wastes like
banana stalk (BS), banana peel (BP), vegetable waste (VW), spent tea waste
(TW) and food waste (FW) in different combinations can be utilized to produce
methane in lab scale fabricated digesters of one litre capacity. The combination of
vegetable waste (VW) and spent tea waste (TW) has produced a maximum biogas
yield of 3.75 L.
Viswanath et al. (1992) carried out a study on anaerobic digestion of fruit and
vegetable wastes for biogas production in a 60 L digester. A maximum biogas yield
of 0.6 m3/kg VS was obtained at a 20 day HRT and 40 kg TS/m3/day organic loading
rate. The biogas production was observed in the digesters operated at 16 and 24 day
HRT. The maximum biogas yield of 74.5% was observed within 12 h of feeding at a
16 day HRT, whereas at a 24 day HRT only 59.03% of biogas yield on substrate
could be generated. Kitchen waste (KW) and fruit/vegetable waste (FVW) could be a
potential candidate for methane production (Wang et al. 2014). The results of
lab-scale experiment showed a higher methane productivity of 0.725 L CH4/g
VS. Similarly, during the anaerobic digestion process of FVW in a batch reactor,
514 ( 57) L CH4/kg VS was produced (Park et al. 2012). Dinsdale et al. (2000)
studied two-stage anaerobic co-digestion of waste activated sludge and FVW using
inclined tubular digesters. The design system could achieve a high biogas yield of
0.37 m3/kg VS. The methane content in the biogas was 68% and bicarbonate
alkalinity in the methanogenic stage was over 4000 mg CaCO3/L.
Manure produced from dairy cows are primarily stored in holding tanks before being
applied to fields in rural areas. This leads to methane emissions in the environment as
the manure decomposes and also may contribute to excess nutrients release in
waterbodies. Biogas production from livestock manure through AD can minimise
such problems, and can also prevent other associated issues of odours and manure
pathogens.
15 Life Cycle Assessment of Anaerobic Digestion Systems: An Approach Towards. . . 401
Most of the rural areas of Asia is still facing problems of large quantities of
manure from dairy, pigs and poultry and the suitable treatment and management of
these waste is required (Vu et al. 2015). By utilizing these large wastes for energy
generation in the form of biogas will help to prevent direct discharge of animal
wastes into aquatic bodies, a common practice in various parts of the Asian region
that lack proper waste management systems (Anenberg et al. 2013; Vu et al. 2015).
Khalil et al. (2019) have reviewed the potential of sustainable biogas production
from animal waste in Indonesia. It was found that about 9597.4 Mm3/year of biogas
can be produced from animal waste in Indonesia and this will generate an electric
power up to 1.7 106 KWh/year.
Achinas and Euverink (2019) have reviewed the biogas production from the
anaerobic co-digestion of farmhouse waste and have studied their performance and
kinetics. El-Mashad and Zhang (2010) have investigated the effect of manure-
screening on the biogas yield from dairy manure in batch digesters under mesophilic
(35 C) conditions. The production of methane from the fine and coarse fraction of
screened manure were 302 and 228 L/kg VS, respectively, and for unscreened
manure it was 241 L/kg VS. Mähnert and Linke (2009) have studied biogas
production from whole-crop rye silage, maize silage and fodder beet silage along
with cattle slurry at mesophilic temperatures. The maximum biogas production and
the rate of biogas generation were in the range of 0.61–0.93 m3/kg volatile solids and
0.032–0.316 per day, respectively. Abubakar and Nasir (2012) have investigated the
effectiveness of cow dung for biogas production using a 10 L volume bioreactor in
batch and semi-continuous mode at mesophilic (53 C) temperature conditions. The
average cumulative biogas yield and methane content were 0.15 L/kg VS and 47%,
respectively. Their study showed that cow dung is an effective feedstock for biogas
production with stable performance.
The AD process of animal manure with other biomass increases the yield of
biogas and it provides advantages for the proper management of manure and such
organic wastes simultaneously (Nielsen et al. 2002). Biogas production from manure
and digestible organic wastes coming from food industry can be an example (Braun
and Wellinger 2003). Uzodinma and Ofoefule (2009) have studied biogas produc-
tion from cow dung, poultry dung, swine dung and rabbit dung blended with field
grass in a 50 L capacity digester. On blending with rabbit dung, cow dung, swine
dung and poultry yielded an average of 7.73 2.86, 7.53 3.84, 5.66 3.77 and
5.07 3.45 L of biogas per total mass of slurry, respectively. The biogas yield
improved during codigestion due to improvement in nutritional quality.
Ojolo et al. (2007) investigated biogas production from poultry wastes, cow dung
and kitchen waste in a 9 L volume reactor. The highest production of biogas was
observed on the 14th day with 85 10–3 dm3 of biogas. The biogas obtained from
poultry droppings, cow dung and kitchen waste were 0.0318 dm3/day, 0.0230 dm3/
day and 0.0143 dm3/day, respectively. In another study, Zhang et al. (2013) found
high biogas production from goat manure by adding with three crops residues at
35 C. The highest cumulative biogas production was 16,023 mL for different
co-substrates after 55 days of digestion.
402 M. M. Hanafiah et al.
Crop residues include materials such as stalks, straw, and plant trimmings, which
needs further processing to dispose of after the cultivation. In general, the crop
residue can be left behind in the fields by which the organic and moisture content of
the soil can be improved. However, where multiple crop cultivations are carried out
successively, this method can cause hindrance to subsequent crops. Furthermore,
higher crop yields leads to higher amounts of residue that need to be taken care of,
creating more problem for the cultivators. Thus, converting such crop residues into
biogas promotes better environmental sustainability. Crop residues are usually
co-digested with other organic materials as it contains a high lignin content that
makes it difficult to break down. However, growing energy crops for the sole
purpose of bioenergy resources comes with a number of downsides such as require-
ment of high costs and large land use, water, and nutrient requirements. In this
context, the biogas obtained from wastes such as crop residues is more
advantageous.
Several agricultural residues in the form of agricultural biomass can be
categorised as the food based portion (oil and simple carbohydrates) of crops
(such as beets, corn and sugarcane) and the non-food waste (complex carbohydrates)
portion (such as rice husk, leaves, orchard trimmings, stalks, corn stover, wheat
straw and pearl millet stalk), perennial grasses and biogenic waste. Rice, maize, and
wheat are the world top three growing crops and their residues are potentially being
used as a good substrate for biofuel production (Sims 2004; Chandra et al. 2012).
Kumar et al. (2018) have reviewed the potential of some cereal crop residues in
India for enhanced biogas production via chemical pre-treatment. They have mainly
focused on chemical pre-treatment of cereal crop residues to generate biogas by
removing lignin. Li et al. (2019b) have evaluated the anaerobic digestion of vege-
table crop residues and their biochemical methane potential. Svensson et al. (2005)
have studied the financial feasibility and prospects of biogas production from crop
residues on a farm-scale level. The study revealed that the high-solids single-stage
fed-batch operation provides the best option for high biogas yield. The methane
yield, operational costs and degree of gas utilization had the strongest impact on the
financial success of the process. Kalra and Panwar (1986) have investigated husk
and straw of rice crops for the AD process in 190 L capacity digesters. 1 kg of rice
straw has yielded about 220 L of biogas during the batch digestion process. A ratio
of rice straw and cattle dung (1:1 dry weight basis) produced 9.1% more gas as
compared to rice straw alone.
In a study using a mixture of agricultural wastes with semi-solid chicken manure,
a maximum methane production of 502 and 506 mL/g VS obtained at 55 C and
35 C, respectively. However, an additional 42% increase in methane production
(695 mL/g VS) was also observed with pretreated (ammonia stripped) chicken
manure compared to other control (untreated) systems (Fatma et al. 2014). The
anaerobic co-digestion of swine manure mixed with crop residues produced 3.5 L/
day of biogas (Cuetos et al. 2011). The semi-continuous experiment was conducted
15 Life Cycle Assessment of Anaerobic Digestion Systems: An Approach Towards. . . 403
in mesophilic reactors having 3 L working capacity for 30 days. Elena et al. (2012)
showed the production of biogas from a mixture of cattle slurry and cheese whey as
the substrate with a maximum biogas production up to 79%.
Life cycle assessment (LCA) is an important decision making tool for determining
alternative methods to reduce the environmental impacts of any process and its
products (Aziz and Hanafiah 2020). Sustainability is the main priority when it comes
to LCA as this methodology suggests the most eco-friendly solution throughout the
process and always has room for improvement to reduce the environmental impacts
(Ismail and Hanafiah 2019a). LCA can help organizations and industries to evaluate
and understand the holistic environmental performance of their products or activities
depending on their sustainability goals towards product improvement, product
innovation or strategic marketing and business plan (Aziz et al. 2019). It allows
organizations and industries to involve in a decision-making process for
implementing and achieving sustainable development and green solutions and
adopt life cycle thinking into their business. Incorporating LCA into their practice
provides the best design and technology options for their products and also ensure
that it does not cause any adverse environmental impact (Ismail and Hanafiah 2020).
Based on ISO 14040 series (ISO 2006a, b), there are four phases in the framework of
the LCA as shown in Fig. 15.3.
The first phase is to identify the goals and scope of a process. During this phase,
the type of product or service, functional units and system boundaries are identified.
Functional units (FU) are the quantity of products to be studied while the boundaries
of the system provide an overview to explain the processes covered in the evaluation.
A functional unit is one the most important aspects in modelling or constructing a
system of a product in LCA. FU is the measurement of a product that describes its
function and it is used in all the calculations throughout the assessment. Products
must have the same functional unit to make comparisons. A product’s feature can
vary from performance, aesthetics, technical quality and additional services up to
costs (Arzoumanidis et al. 2020). For instance, a material such as food waste can be
described as 1 tonne of food waste, whereas energy can be described as 1 kWh of
electricity. System boundaries determine the inputs that should be included or
excluded in the LCA assessment.
The second phase is the analysis of the life cycle inventory (LCI). It is a part of a
LCA where all the data of a system are collected. It is a complex portion of an
assessment as it tracks and analyses the entire flow of the input and output data of a
404 M. M. Hanafiah et al.
Interpretaon
product’s system. LCI can include material use, energy use, coproduct generation,
waste generation, raw material extraction, refining and processing, product
manufacturing, product use, and recycling or disposal of a product in a system
(DeRosa and Allen 2017)
The third phase in the LCA framework is the life cycle impact assessment
(LCIA). The environmental impact potential is identified using the results of the
life cycle inventory analysis that will be sorted into major impact categories which
are divided into three different groups: ecosystem impacts; human impacts; and
resource depletion. These categories will subsequently be assessed for decision
making. The fourth phase is the interpretation. In this phase, the results obtained
from the previous phases are interpreted in an informative form. In addition,
evaluations and recommendations to reduce the environmental impact of the pro-
duction of a product or service are also reported. The four phases in this life cycle
evaluation framework demonstrate a systematic process of providing relative and
quantitative information to measure the environmental load at each stage of the
product’s life cycle.
As illustrated in Fig. 15.4, LCA uses cradle-to-cradle or cradle-to-grave tech-
niques which analyse raw material extraction through materials processing, manu-
facture, distribution, product usage and disposal or end of life.
2021; Harun et al. 2020; Ismail and Hanafiah 2019b). Below are the historical
background of the introduction and application of LCA for products and systems.
(a) 1960–1990: decades of genesis
The initial studies that are now recognised as partial LCAs, date from the late
1960s to early 1970s. The scope of study during this era was limited to energy
analyses but gradually broadened to encompass resource requirements, emission
loadings and generated waste. LCA mainly focused on covering the alternatives
in this period. The first impact assessment method was introduced in the early
1980s, separating the airborne and waterborne emissions by semi-political
standards for those emissions and aggregating them into ‘critical volumes’ of
air and ‘critical volumes’ of water, respectively. From 1970s to 1980s, there was
a lack of international scientific discussion and information exchange on LCA
platforms. Thus, LCAs were performed using different methods and without any
common standard structure. This prevented LCA to be accepted more generally
as an analytic tool as there were great differences in the results obtained even
when the same materials were used (Guinee 2011).
(b) 1990–2000: decade of standardisation
The tremendous elevation in scientific and coordination activities world-wide
reflected in the increasing number of LCA guides and handbooks produced. The
International Organisation for Standardisation (ISO) has been involved in LCA
since 1994. Other than that, the first scientific journal papers began to appear in
the Journal of Cleaner Production; Resources, Conservation and Recycling;
International Journal of LCA; Environmental Science and Technology; Indus-
trial Ecology and other journals. Moreover, LCA became part of policy docu-
ments and legislation during these years. Several well-recognised impact
assessment methods such as CML 1992 environmental theme approach that
are still being used till today were developed in this period.
(c) 2000–till now: decades of expansion
The attention towards LCA began to grow in the first decade of the twenty-
first century. In 2002, the United Nations Environment Programme (UNEP) and
the society for Environmental Toxicology and Chemistry (SETAC) launched an
International Life Cycle Partnership known as the Life Cycle Initiative, where
their main aim was formulated as putting life cycle thinking into practice and
improving the supporting tools through better data and indicators. The European
406 M. M. Hanafiah et al.
Figure 15.5 illustrates the phases involved in the LCA of AD system. Some previous
studies on life cycle assessment of anaerobic digestion system are given in
Table 15.2.
Mezzullo et al. (2013) evaluated the environmental impacts of biogas production
and utilization through the framework of LCA. The production of biogas and
fertilizer (from digestate) from a cattle farm waste is compared with other alternative
forms of energy including fossil fuel. The biogas based system was much more
advantages in terms of greenhouse gas emission and fossil-fuel usage. In terms of
Table 15.2 Selected studies on life cycle assessment of anaerobic digestion process
Organic waste Impact
Location source Functional unit Software method Reference
Australia Food waste 1 kg of food SinaPro ReCiPe Opatokun et al.
waste v.8 (2017)
Italy Agricultural prod- 1 MWh of Gabi CML Fusi et al.
ucts and waste electricity LCA 2001 (2016)
v6.11
United Cattle waste 1 m3 of biogas SimaPro EI 99 Mezullo et al.
Kingdom (2013)
United Cow manure 1 kg of excreted GaBi 5 IPCC, Andres and
States manure 2006 Rebecca (2019)
Vietnam Pig manure 100 kg of solid ReCiPe, Vu et al. (2015)
pig manure 2008
1000 kg of liquid
pig manure
Norway Fuel 1 kg of CO2 SimaPro CMLIAb Lyng and
v8.5.2 v3.05 Brekke (2019)
IPCC
(2013)
Malaysia Palm oil mill 1 tonne of POME SimaPro ReCiPe, Aziz and
effluent (POME) v8.5 2016 Hanafiah (2020)
Italy Organic waste 1 tonne of CML Di Maria and
organic waste Micale (2015)
both environmental and energy impact, the system contributed negligible amounts
towards the whole life cycle impacts.
Fusi et al. (2016) evaluated the environmental impacts (through LCA) associated
with generation of electricity from biogas produced by AD of agricultural wastes in
five actual plants. The results suggest that the main contributors to the impacts are
the use of maize silage, the operation of the anaerobic digester, including open
storage of digestate. The small scale system using animal slurry is the best option
among the different plants considered in this study. The only impacts an animal
slurry based biogas plants has are marine and terrestrial ecotoxicity. However, in
comparison to other renewable sources and natural gases for electricity generation,
biogas based electricity performed poorly with higher environmental impacts.
According to the author, the environmental impacts could be mitigated by under-
taking better practices such as avoiding digestate storage, preventing biogas emis-
sions and reducing digestate application to land.
Vu et al. (2015) reported that loss of biogas from digesters as well as the
intentional release of biogas and emissions of CH4 from manure storage compromise
the beneficial effects of biogas use. However, from the sensitivity analysis, it is also
clear that biogas digesters can become a means of reducing global warming impacts
relatively easy if CH4 emissions into the environment can be kept low. In another
study, Lyng and Brekke (2019) stated that when applying life cycle assessment to
evaluate the environmental impacts from biogas as a fuel for transport, the results are
408 M. M. Hanafiah et al.
Most countries deal with the wastes according to their respective waste hierarchy.
Although, the steps of the waste hierarchy of a country may slightly vary from one
another, its ultimate motive is to work towards sustainable waste management. In
general, the first step of a waste hierarchy is to reduce waste generation at the point of
source. Reducing waste needs to be considered from the beginning of designing any
sort of product in order to minimise the toxicity and the amount of waste generated
throughout the product’s life. This step emphasises the importance to carry out a
LCA before manufacturing a product. Apart from that, reducing waste can be
practiced in every household as this is the simplest way to curb problems related
to waste management. However, this step might have a drawback when it comes to
changing from citizens’ up to stakeholders’ behaviour and attitude towards sustain-
able waste management as this step can only be completely effectively if everyone
joins forces to do their responsibility in reducing waste.
The second step of the waste hierarchy is to reuse a product more than once.
Accumulation of the amount of waste in the landfill can be decreased by practicing
this step. This step also requires a strong commitment from the majority of the
people to succeed. Recycling is the third step in the hierarchy of waste. Similar to the
two other steps mentioned before, recycling also starts from home by segregating
waste according to its type. Recycling is a very resourceful step as it extracts the raw
material from a product to be reused or to generate new products and reduce
15 Life Cycle Assessment of Anaerobic Digestion Systems: An Approach Towards. . . 409
landfilling as well. Anaerobic digestion and composting are included in this step as it
recovers useful organic compounds from biodegradable waste.
Although composting is a traditional method that recovers and dissolves organic
material back to the land, it is considered one the most effective methods as it is
simple to carry out and the cost is minimal. However, this method may attract pests
in an area. The next level of this hierarchy is the treatment or energy recovery step,
whereby energy is recovered from waste through various types of treatment or
recovery techniques. Anaerobic digestion can also be included in this category as
it generates biogas that can be used as a source of energy to produce electricity and
fuel as discussed in this chapter. Biological anaerobic digestion gains more impor-
tance among the various conversion processes of biomass due to its economical and
efficient way of recovering carbon in the form of renewable biogas fuel (Achinas
et al. 2020). Other recovery techniques include mass burn, pyrolysis and refuse-
derived fuel (RDF). Lastly, the bottom of the hierarchy is categorised as landfilling
because it causes many environmental issues such as pollution and global warming
due to high emissions of GHGs, leachate, and toxicants.
References
Abubakar BSUI, Nasir I (2012) Anaerobic digestion of cow dung for biogas production. ARPN J
Eng Appl Sci 7(2)
Abouelenien F, Namba Y, Kosseva MR, Nishio N, Nakashimada Y (2014) Enhancement of
methane production from co-digestion of chicken manure with agricultural wastes. Bioresour
Technol 159:80–87
Achinas S, Euverink GJW (2019) Elevated biogas production from the anaerobic co-digestion of
farmhouse waste: insight into the process performance and kinetics. Waste Manag Res 37
(12):1240–1249
Achinas S, Achinas V, Euverink GJW (2020) Chapter 2—Microbiology and biochemistry of
anaerobic digesters: an overview. Bioreactors 17–26. https://www.sciencedirect.com/science/
article/pii/B9780128212646000024
Akindele AA, Sartaj M (2017) The toxicity effects of ammonia on anaerobic digestion of organic
fraction of municipal solid waste. Waste Manag 71:757–766
Al-Raad AA, Hanafiah MM, Naje AS, Ajeel MA (2020) Optimized parameters of the
electrocoagulation process using a novel reactor with rotating anode for saline water treatment.
Environ Pollut 265:115049
Al-Wahaibi A, Osman AI, Al-Muhtase A’a H, Alqaisi O, Baawain M, Fawzy S, Rooney DW (2020)
Techno economic evaluation of biogas production from food waste via anaerobic digestion. Sci
Rep 10:15719
Andres AH, Rebecca L (2019) From waste to energy: life cycle assessment of anaerobic digestion
systems. https://lpelc.org/from-waste-to-energy-life-cycle-assessment-of-anaerobic-digestion-
systems/
Anenberg SC, Balakrishnan K, Jetter J, Masera O, Mehta S, Moss J, Ramanathan V (2013) Cleaner
cooking solutions to achieve health, climate, and economic co-benefits. Environ Sci Technol 47
(9):3944–3952
410 M. M. Hanafiah et al.
Appels L, Lauwers J, Degrève J, Helsen L, Lievens B, Willems JVI, Dewil R (2011) Anaerobic
digestion in global bio-energy production: potential and research challenges. Renew Sust Energ
Rev 15:4295–4301
Arhoun B, Villen-Guzman MD, Vereda-Alonso C, Rodriguez-Maroto JM, Garcia-Herruzo F,
Gomez-Lahoz C (2019) Anaerobic co-digestion of municipal sewage sludge and fruit/vegetable
waste: effect of different mixtures on digester stability and methane yield. J Environ Sci Health
A Tox Hazard Subst Environ Eng 1:1–7
Arzoumanidis I, D’Eusanio M, Raggi A, Petti L (2020) Functional unit definition criteria in life
cycle assessment and social life cycle assessment: a discussion. In: Traverso M, Petti L,
Zamagni A (eds) Perspectives on social LCA. SpringerBriefs in environmental science.
Springer, Cham. https://doi.org/10.1007/978-3-030-01508-4_1
Ashraf MA, Hanafiah MM (2017) Recent advances in assessment on clear water, soil and air.
Environ Sci Pollut Res 24(29):22753–22754
Ashraf MA, Hanafiah MM (2019) Sustaining life on earth system through clean air, pure water, and
fertile soil. Environ Sci Pollut Res 26:13679–13680
Ashraf MA, Balkhair KS, Chowdhury AJK, Hanafiah MM (2019) Treatment of Taman Beringin
landfill leachate using the column technique. Desalin Water Treat 149:370–387
Aziz NIHA, Hanafiah MM, Gheewala SH, Ismail H (2020) Bioenergy for a cleaner future: a case
study of sustainable biogas supply chain in the Malaysian energy sector. Sustainability 12:3213
Aziz NIHA, Hanafiah MM (2021) Application of life cycle assessment for desalination: progress,
challenges and future directions. Environ Pollut 268(Part B):115948
Aziz NIHA, Hanafiah MM, Gheewala SH (2019) A review on life cycle assessment of biogas
production: challenges and future perspectives in Malaysia. Biomass Bioenergy 122:361–374
Banch TJ, Hanafiah MM, Alkarkhi AF, Amr A, Salem S (2019) Factorial design and optimization
of landfill leachate treatment using tannin-based natural coagulant. Polymers 11:1349
Banch TJ, Hanafiah MM, Alkarkhi AF, Amr A, Amr SSA, Nizam NUM (2020a) Evaluation of
different treatment processes for landfill leachate using low-cost agro-industrial materials.
Processes 8(11):1–12
Banch TJ, Hanafiah MM, Amr SSA, Alkarkhi AF, Amr A, Hasan M (2020b) Treatment of landfill
leachate using palm oil mill effluent. Processes 8(601):1–17
Bouallagui H, Lahdheb H, Romdan EB, Rachdi B, Hamdi M (2009) Improvement of fruit and
vegetable waste anaerobic digestion performance and stability with co-substrates addition. J
Environ Manag 90(5):1844–1849
Braun R, Wellinger A (2003) Potential of co-digestion. IEA Bioenergy, task 37—energy from
biogas and landfill gas. www.IEA-Biogas.net
Browne J, Murphy J (2013) Assessment of the resource associated with biomethane from food
waste. Appl Energy 104:170–177
Carlsson M, Naroznova I, Moller J, Scheutz C, Lagerkvist A (2015) Importance of food waste
pre-treatment efficiency for global warming potential in life cycle assessment of anaerobic
digestion systems. Resour Conserv Recycl 102:58–66
Chanakya HN, Ramachandra TV, Vijayachamundeeswar M (2007) Resource recovery potential
from secondary components of segregated municipal solid wastes. Environ Monit Assess
135:119–127
Chandra R, Takeuchi H, Hasegawa T (2012) Methane production from lignocellulosic agricultural
crop wastes: a review in context to second generation of biofuel production. Renew Sust Energ
Rev 16(3):1462–1476
Chukwudi O, Ifeanyichukwu E, Victor I, Nneka O (2019) Towards effective management of
digester dysfunction during anaerobic treatment processes. Renew Sust Energ Rev 116:109424
Comino E, Riggio VA, Rosso M (2012) Biogas production by anaerobic co-digestion of cattle
slurry and cheese whey. Bioresour Technology 114:46–53
Cuetos MJ, Fernández C, Gómez X, Morán A (2011) Anaerobic co-digestion of swine manure with
energy crop residues. Biotechnol Bioprocess Eng 16:1044–1052
15 Life Cycle Assessment of Anaerobic Digestion Systems: An Approach Towards. . . 411
DeRosa SE, Allen DT (2017) Comparison of attributional and consequential life-cycle assessments
in chemical manufacturing. Encyclopedia Sustain Technol:339–347
Di Maria F, Micale C (2015) Life cycle analysis of incineration compared to anaerobic digestion
followed by composting for managing organic waste: the influence of system components for an
Italian district. Int J Life Cycle Assess 20:377–388
Dinsdale RM, Premier GC, Hawkes FR, Hawkes DL (2000) Two-stage anaerobic co-digestion of
waste activated sludge and fruit/vegetable waste using inclined tubular digesters. Bioresour
Technol 72:159–168
El-Mashad HM, Zhang R (2010) Biogas production from co-digestion of dairy manure and food
waste. Bioresour Technol 101:4021–4028
FAO (2011) Global food losses and food waste. http://www.fao.org/docrep/014/mb060e/
mb060e00.pdf
Farhat A, Miladi B, Hamdi M, Bouallagui H (2018) Fermentative hydrogen and methane
co-production from anaerobic co-digestion of organic wastes at high loading rate coupling
continuously and sequencing batch digesters. Environ Sci Pollut Res 25(28):27945–27958
Fatma A, Yuzaburo N, Maria RK, Naomichi N, Yutaka N. (2014) Enhancement of methane
production from co-digestion of chicken manure with agricultural wastes. Bioresour Technol
159; 80–87.
Fusi A, Bacenetti J, Fiala M, Azapagic A (2016) Life cycle environmental impacts of electricity
from biogas produced by anaerobic digestion. Front Bioeng Biotechnol 4:26
Gaby JC, Zamanzadeh M, Horn SJ (2017) The effect of temperature and retention time on methane
production and microbial community composition in staged anaerobic digesters fed with food
waste. Biotechnol Biofuels 10(1):1–13
Gao M, Zhang S, Ma X, Guan W, Song N, Wang Q, Wu C (2020) Effect of yeast addition on the
biogas production performance of a food waste anaerobic digestion system. R Soc Open Sci
7:200443
Garcia-Peña EI, Parameswaran P, Kang DW, Canul-Chan M, Krajmalnik-Brown R (2011) Anaer-
obic digestion and co-digestion processes of vegetable and fruit residues: process and microbial
ecology. Bioresour Technol 102:9447–9455
García-Pérez J, Fernández-Navarro P, Castelló A, López-Cima MF, Ramis R, Boldo E et al (2013)
Cancer mortality in towns in the vicinity of incinerators and installations for the recovery or
disposal of hazardous waste. Environ Int 51:31–44
González-Sánchez ME, Pérez-Fabiel S, Wong-Villarreal A, Bello-Mendoza R, Yanez-Ocampo G
(2015) Agro-industrial wastes methanization and bacterial composition in anaerobic digestion.
Rev Argent Microbiol 47(3):229–235
Guinee JB (2011) International symposium on life cycle assessment and construction. July 10-12,
Nantes, France. Life cycle assessment: past, present and future. Environ Sci Technol 45:90–96
Gupta P, Singh RS, Sachan A, Vidyarthi AS, Gupta A, Tripathi N (2010) Study on biogas
production from vegetable waste by indigenous microbes. Int J Appl Environ Sci 5(2)
Gupta P, Singh RS, Sachan A, Vidyarthi AS, Gupta A (2012a) Study on biogas production by
anaerobic digestion of garden-waste. Fuel 95:495–498
Gupta P, Singh RS, Sachan A, Vidyarthi AS, Gupta A (2012b) A re-appraisal on intensification of
biogas production. Renew Sust Energ Rev 16:4908–4916
Hanafiah MM, Hashim NA, Ahmed ST, Ashraf MA (2018) Removal of chromium from aqueous
solutions using a palm kernel shell adsorbent. Desalin Water Treat 118:172–180
Hanafiah MM, Zainuddin MF, Mohd Nizam NU, Halim AA, Rasool A (2020) Phytoremediation of
aluminium and iron from industrial wastewater using Ipomoea aquatica and Centella asiatica.
Appl Sci 10:3064
Harun SN, Hanafiah MM, Aziz NIHA (2020) An LCA-based environmental performance of rice
production for developing a sustainable agri-food system in Malaysia. Environ Manag 67
(1):146–161. https://doi.org/10.1007/s00267-020-01365-7
Hoo PY, Hashim H, Ho WS, Tan ST (2017) Potential biogas generation from food waste through
anaerobic digestion in Peninsular Malaysia. Chem Eng Trans 56:373–378
412 M. M. Hanafiah et al.
Hundal JS, Wadhwa M, Bakshi MPS (2019) Herbal feed additives containing essential oil: impact
on the nutritional worth of complete feed in vitro. Trop Anim Health Prod 51:1909–1917
International Organization for Standardization (ISO) (2006a) ISO 14040:2006 environmental
management—life cycle assessment—principles and framework
International Organization for Standardization (ISO) (2006b) ISO 14044:2006 environmental
management—life cycle assessment—requirements and guidelines
Ismail H, Hanafiah MM (2019a) An overview of LCA application in WEEE management: current
practices, progress and challenges. J Clean Prod 232:79–93
Ismail H, Hanafiah MM (2019b) Discovering opportunities to meet the challenges of an effective
waste electrical and electronic equipment recycling system in Malaysia. J Clean Prod
238:117927
Ismail H, Hanafiah MM (2020) A review of sustainable e-waste generation and management:
present and future perspectives. J Environ Manag 264:110495
Jeong-Ik O, Lee J, Lin K-YA, Kwon EE, Tsang YF (2018) Biogas production from food waste via
anaerobic digestion with wood chips. Energy Environ:1–8
Kalra MS, Panwar JS (1986) Anaerobic digestion of rice crop residues. Agric Wastes 17:263–269
Khalil M, Berawi MA, Heryanto R, Rizalie A (2019) Waste to energy technology: the potential of
sustainable biogas production from animal waste in Indonesia. Renew Sust Energ Rev
105:323–331
Kougias PG, Angelidaki I (2018) Biogas and its opportunities—a review keywords. Front Environ
Sci 12:1–22
Kumar S, Paritosh K, Pareek N, Chawade A, Vivekanand V (2018) De-construction of major Indian
cereal crop residues through chemical pretreatment for improved biogas production: an over-
view. Renew Sust Energ Rev 90:160–170
Li Y, Jin Y, Borrion A, Li H, Li J (2017) Effects of organic composition on the anaerobic
biodegradability of food waste. Bioresour Technol 243:836–845
Li Y, Chen Y, Wu J (2019a) Enhancement of methane production in anaerobic digestion process: a
review. Appl Energy 240:120–137
Li P, Li W, Sun M, Xu X, Zhang B, Sun Y (2019b) Evaluation of biochemical methane potential
and kinetics on the anaerobic digestion of vegetable crop residues. Energies 12:26
Lin J, Zuo J, Gan L, Li P, Liu F, Wang K, Chen L, Gan H (2011) Effects of mixture ratio on
anaerobic co-digestion with fruit and vegetable waste and food waste of China. J Environ Sci
(China) 23(8):1403–1408
Lin C, Pfaltzgraff L, Herrero-Davila L, Mubofu E, Abderrahim S, Clark J, Koutinas A,
Kopsahelis N, Stamatelatou K, Dickson F, Thankappan S, Mohamed Z, Brocklesby R, Luque
R (2013) Food waste as a valuable resource for the production of chemicals, materials and fuels.
Current situation and global perspective. Energy Environ Sci 6:426–464
Liu X, Gao X, Wang W, Zheng L, Zhou Y, Sun Y (2012) Pilot-scale anaerobic co-digestion of
municipal biomass waste: focusing on biogas production and GHG reduction. Renew Energy
44:463–468
Lyng K, Brekke A (2019) Environmental life cycle assessment of biogas as a fuel for transport
compared with alternative fuels. Energies 12:532
Mähnert P, Linke B (2009) Kinetic study of biogas production from energy crops and animal waste
slurry: effect of organic loading rate and reactor size. Environ Technol 30(1):93–99
Maile I, Muzenda E, Mbohwa C (2016) Biogas production from anaerobic digestion of fruit and
vegetable waste from Johannesburg market. In: Proceedings of 2016 international conference on
San Francisco
Mao C, Feng Y, Wang X, Ren G (2015) Review on research achievements of biogas from anaerobic
digestion. Renew Sust Energ Rev 45:540–555
Masebinu SO, Akinlabi ET, Muzenda E, Aboyade AO, Mbohwa C (2018) Experimental and
feasibility assessment of biogas production by anaerobic digestion of fruit and vegetable
waste from Joburg Market. Waste Manag 75:236–250
15 Life Cycle Assessment of Anaerobic Digestion Systems: An Approach Towards. . . 413
Mezzullo WG, Mcmanus MC, Hammond GP (2013) Life cycle assessment of a small-scale
anaerobic digestion plant from cattle waste. Appl Energy 102:657–664
Morales-Polo C, Cledera-Castro MDM, Soria MBY (2018) Reviewing the anaerobic digestion of
food waste: from waste generation and anaerobic process to its perspectives. Appl Sci 8
(10):1804. https://doi.org/10.3390/app8101804
Nielsen LH, Hjort-Gregersen K, Thygesen P, Christensen J (2002) Samfundsøkonomiske analyser
af biogasf_llesanl_g. Rapport 136. Fødevareøkonomisk Institut, København (Summary in
English)
Nizam NUM, Mohd Hanafiah M, Mohd Noor I, Abd Karim HI (2020) Efficiency of five selected
aquatic plants in phytoremediation of aquaculture wastewater. Appl Sci 10:2712
Ojolo SJ, Oke SA, Animasahun K, Adesuyi BK (2007) Utilization of poultry, cow and kitchen
wastes for biogas production: a comparative analysis Iran. J Environ Health Sci Eng 4
(4):223–228
Opatokun SA, Ana M, Ferreira M, Strezov V (2017) Life cycle analysis of energy production from
food waste through anaerobic digestion, pyrolysis and integrated energy system. Sustainability
9:1804
Pallan AP, Antony Raja S, Varma CG, Deepak Mathew DK, Anil KS, Kannan A (2018) Biogas
production from community waste to optimise the substrate for anaerobic digestion. Int Rev
Mech Eng (IREME) 12(7):580–589
Papargyropoulou E, Lozano R, Steinberger JK, Wright N, Ujang ZB (2014) The food waste
hierarchy as a framework for the management of food surplus and food waste. J Clean Prod
76:106–115
Paritosh K, Kushwaha SK, Yadav M, Pareek N, Chawade A, Vivekanand V (2017) Food waste to
energy: an overview of sustainable approaches for food waste management and nutrient
recycling. BioMed Research Int 19. https://doi.org/10.1155/2017/2370927
Park ND, Thring RW, Helle SS (2012) Comparison of methane production by co-digesting fruit and
vegetable waste with first stage and second stage anaerobic digester sludge from a two stage
digester. Water Sci Technol 65(7):1252–1257
Pavi S, Kramer LE, Gomes LP, Miranda LAS (2017) Biogas production from co-digestion of
organic fraction of municipal solid waste and fruit and vegetable waste. Bioresour Technol
228:362–367
Ranade DR, Yeole TY, Godbole SH (1987) Production of biogas from market waste. Biomass
13:147–153
Ren Y, Yu M, Wu C, Wang Q, Gao M, Huang Q, Liu Y (2018) A comprehensive review on food
waste anaerobic digestion: research updates and tendencies. Bioresour Technol 247:1069–1076
Saraswat M, Garg M, Bhardwaj M, Mehrotra M, Singhal R (2019) Impact of variables affecting
biogas production from biomass. IOP Conf Ser Mater Sci Eng 691
Scano EA, Asquer C, Pistis A, Ortu L, Demontis V, Cocco D (2014) Biogas from anaerobic
digestion of fruit and vegetable wastes: experimental results on pilot-scale and preliminary
performance evaluation of a full-scale power plant. Energy Convers Manag 77:22–30
Sims RE (2004) Biomass, bioenergy and biomaterials—future prospects. Biomass and agricul-
ture—sustainability markets and policies. OECD, Paris, pp 37–61
Sitorus B, Sukandar S, Panjaitan SD (2013) Biogas recovery from anaerobic digestion process of
mixed fruit–vegetable wastes. Energy Proc 32:176–182
Svensson LM, Christensson K, Bjornsson L (2005) Biogas production from crop residues on a
farm-scale level: is it economically feasible under conditions in Sweden? Bioprocess Biosyst
Eng 28:139–148
Uzodinma EO, Ofoefule AU (2009) Biogas production from blends of field grass (Panicum
maximum) with some animal wastes. Int J Phys Sci 4(2):091–095
Vambol S, Vambol V, Sundararajan M, Ansari I (2019) The nature and detection of unauthorized
waste dump sites using remote sensing. Ecol Quest 30(3)
Viswanath P, Devi SS, Nand K (1992) Anaerobic digestion of fruit and vegetable processing wastes
for biogas production. Bioresour Technol 40(1):43–48
414 M. M. Hanafiah et al.
Vu TKV, Vu DQ, Jensen LS, Sommer SG, Bruun S (2015) Life cycle assessment of biogas
production in small-scale household digesters in Vietnam. Asian-Australas J Anim Sci 28
(5):716–729
Wang L, Shen F, Yuan H, Zou D, Liu Y, Zhu B, Li X (2014) Anaerobic co-digestion of kitchen
waste and fruit/vegetable waste: lab-scale and pilot-scale studies. Waste Manag 34
(12):2627–2633
Winichayakul S et al (2020) In vitro gas production and rumen fermentation profile of fresh and
ensiled genetically modified high-metabolizable energy ryegrass. J Dairy Sci 103:2405–2418
Wu Y, Wang C, Liu X, Ma H, Wu J, Zuo J, Wang K (2016) A new method of two-phase anaerobic
digestion for fruit and vegetable waste treatment. Bioresour Technol 21:16–23
Yang Y-Q, Shen D-S, Li N, Xu D, Long Y-Y, Lu X-Y (2013) Co-digestion of kitchen waste and
fruit–vegetable waste by two-phase anaerobic digestion. Environ Sci Pollut Res 20:2162–2171
Yang G, Zhang P, Zhang G, Wang Y, Yang A (2015) Degradation properties of protein and
carbohydrate during sludge anaerobic digestion. Bioresour Technol 192:126–130
Yazdanpanah A, Ghasimi DSM, Kim MG, Nakhla G, Hafez H, Keleman M (2018) Impact of trace
element supplementation on mesophilic anaerobic digestion of food waste using Fe-rich inoc-
ulum. Environ Sci Pollut Res Int 25(29):29240–29255
Zhang T, Liu L, Song Z, Ren G, Feng Y et al (2013) Biogas Production by co-digestion of goat
manure with three crop residues. PLoS One 8(6):e66845