Download as pdf or txt
Download as pdf or txt
You are on page 1of 334

Arun Lal Srivastav

Abhishek Kumar Bhardwaj


Mukesh Kumar Editors

Valorization
of Biomass Wastes
for Environmental
Sustainability
Green Practices for the Rural Circular
Economy
Valorization of Biomass Wastes for Environmental
Sustainability
Arun Lal Srivastav • Abhishek Kumar Bhardwaj
Mukesh Kumar
Editors

Valorization of Biomass
Wastes for Environmental
Sustainability
Green Practices for the Rural Circular
Economy
Editors
Arun Lal Srivastav Abhishek Kumar Bhardwaj
Chitkara University School of Engineering School of Life Sciences
and Technology Amity University Madhya Pradesh
Chitkara University Gwalior, Madhya Pradesh, India
Solan, Himachal Pradesh, India

Mukesh Kumar
Sardar Vallabhbhai Patel University
of Agriculture and Technology
Meerut, Uttar Pradesh, India

ISBN 978-3-031-52484-4    ISBN 978-3-031-52485-1 (eBook)


https://doi.org/10.1007/978-3-031-52485-1

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Switzerland AG 2024
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

Paper in this product is recyclable.


Contents

 Biomass Energy a Boon or Bane for Society: A Comprehensive


Is
Analysis ������������������������������������������������������������������������������������������������������������    1
Shama E. Haque and Tausif Rahman Rafi
Approach to Reduce Agricultural Waste via Sustainable
Agricultural Practices��������������������������������������������������������������������������������������   21
Prasann Kumar, Amit Raj, and Vantipalli Aravind Kumar
Biomass Waste and Bioenergy Production: Challenges
and Alternatives������������������������������������������������������������������������������������������������   51
Ahmed Albahnasawi, Murat Eyvaz, Motasem Y. D. Alazaiza,
Nurullah Özdoğan, Ercan Gurbulak, Sahar Alhout,
and Ebubekir Yuksel
Enzyme-Mediated Strategies for Effective Management
and Valorization of Biomass Waste����������������������������������������������������������������   69
Usman Lawal Usman, Bharat Kumar Allam, and Sushmita Banerjee
Nanotechnological Advancements for Enhancing Lignocellulosic
Biomass Valorization���������������������������������������������������������������������������������������   99
Vijayalakshmi Ghosh
A State of the Art of Biofuel Production Using Biomass Wastes:
Future Perspectives������������������������������������������������������������������������������������������ 115
Thi An Hang Nguyen, Thi Viet Ha Tran, and Minh Viet Nguyen
Role of Pretreatment Approaches to Generate Value-Added
Products Using Agriculture Biomass�������������������������������������������������������������� 133
Suman, Deepanshu Awasthi, Nishtha, Nikhil Gakkhar, and Bharat Bajaj
Utilising Biomass-Derived Composites in 3D Printing
to Develop Eco-Friendly Environment���������������������������������������������������������� 153
Chetan Chauhan, Varsha Rani, Mukesh Kumar, and Rishubh Motla

v
vi Contents

Bioenergy Production Using Biomass Wastes: Challenges


of Circular Economy���������������������������������������������������������������������������������������� 171
Vijaya Ilango

Application of Enzymes in Biomass Waste Management���������������������������� 189
Preeti Ranjan, Maneesh Kumar, Himanshu Bhardwaj,
Priyanka Kumari, and Arti Kumari
Pretreatment Techniques for Derivation of Value-Added
Products from Agro-Waste Biomass�������������������������������������������������������������� 207
Tran Thi Viet Ha, Nguyen Thi An Hang, and Nguyen Minh Viet
Significance of Enzymatic Actions in Biomass Waste
Management: Challenges and Future Scope ������������������������������������������������ 223
Prangya Rath, Laxmi Kant Bhardwaj, Mini Chaturvedi,
and Abhishek Bhardwaj
Bioeconomy: A Sustainable Approach for Biomass
Waste Management������������������������������������������������������������������������������������������ 239
Rwitabrata Mallick, Kuldip Dwivedi, and Swapnil Rai

Application of Flower Wastes to Produce Valuable Products���������������������� 251
Avnish Chauhan, Manya Chauhan, Muneesh Sethi, Arvind Bodhe,
Anirudh Tomar, Shikha, and Nitesh Singh
Myco-degradation of Lignocellulosic Waste Biomass
and Their Applications������������������������������������������������������������������������������������ 269
Sahith Chepyala, Jagadeesh Bathula, and Sreedhar Bodiga

Value-Added Product Development Utilising the Food Wastes ������������������ 287
Anduri Sravani, C. R. Patil, and Shivani Sharma
Role of Bacterial Degradation in Lignocellulosic Biomass
for Biofuel Production ������������������������������������������������������������������������������������ 303
Arti Kumari, Maneesh Kumar, and Bibekananda Bhoi
Cultivating a Greener Tomorrow: Sustainable Agriculture
Strategies for Minimizing Agricultural Waste���������������������������������������������� 317
Dipti Bharti, Abhilekha Sharma, Meenakshi Sharma, Rahul Singh,
Amit Kumar, and Richa Saxena

Index������������������������������������������������������������������������������������������������������������������ 335
Is Biomass Energy a Boon or Bane
for Society: A Comprehensive Analysis

Shama E. Haque and Tausif Rahman Rafi

1 Introduction

Fossil fuels (e.g., petroleum, coal, and natural gas) have historically dominated the
energy sector. Combustion of fossil fuel is the largest source of greenhouse gas
(GHGs) emissions from human activities that is driving global climate change.
Anthropogenic GHG emissions are altering our planet’s energy balance between
incoming solar radiation and the heat released back into space, increasing the green-
house effect, which is resulting in global climate change (Haque, 2023; Haque &
Nahar, 2023; USEPA, 2023a).Renewable biomass energy, derived from biodegrad-
able fractions of products, natural resources such as waste biomass, and residues of
biological origin from agricultural fields, wood industrialization, municipal prun-
ing, and food processing, has tremendous potential to mitigate global climate
change (Osman et al., 2021; Tripathi et al., 2019; Barbieri et al., 2013) resulting
from the transportation sector (Przywara et al., 2023). There are numerous disposal
concerns and governance issues due to the rapidly increasing amount of biomass
waste (Zhou & Wang, 2020; Tripathi et al., 2019). Specifically, in developing coun-
tries, the majority of biomass residues are either burned outdoors or left in the field
to decay, having a negative impact on the environment. Due to rapid urbanization
and the rising demand for construction products, there is an increasing need for
energy and biomass wastes remains an underutilized resource (Tripathi et al., 2019).
Waste biomass from the agricultural industry can be converted to produce energy
and could serve to be a way to reduce the land occupied by landfills. Biomass con-
version is the process of transforming organic plant matter into fuel and, subse-
quently, utilizing the fuel as a source of energy. Biomass waste can be combusted to
generate heat (direct), converted into electricity (direct), or processed into biofuel

S. E. Haque (*) · T. R. Rafi


North South University, Bashundhara, Dhaka, Bangladesh
e-mail: shama.haque@northsouth.edu

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 1


A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_1
2 S. E. Haque and T. R. Rafi

(indirect). Biofuel production varies according to the types of raw materials, level of
efficiency, volume produced, environmental conditions, and user requirements
(Przywara et al., 2023). A minor fraction of biomass waste generated becomes a
feedstock for industrial applications and electricity generation, and the remaining
negatively influences the atmosphere, surface water, and groundwater quality issues
(Tripathi et al., 2019). The International Energy Agency (IEA, 2011) reports that
biofuels have the potential to supply approximately 27% of global transportation
fuel demand by 2050.
Since the 1990s, the global consumption of primary energy increased dramati-
cally, leading to a shortage of primary energy and increased GHG emissions
(USEPA, 2023a). Since 1970, atmospheric carbon dioxide (CO2) emission has
increased by more than 90%, with emissions from fossil fuel combustion and indus-
trial operations accounting for approximately 78% of the entire increase in GHGs
between 1970 and 2011. Additionally, recently, the lockdown caused by the
COVID-19 pandemic has greatly affected the renewable energy sector, contributing
to a collapse in the price of oil and lower prices for other fossil fuels (Jiang et al.,
2021). However, since then, the world has witnessed a remarkably quick economic
rebound, and in 2021, anthropogenic CO2 emissions from fossil fuel combustion
increased to their highest yearly level ever (IEA, 2022). Subsequently, in 2022,
Russia’s invasion of Ukraine directly affected the costs of heating, cooling, and
lighting and contributed to a global energy crisis (Guan et al., 2023; IEA, 2022).
Due to the volatility of the global energy market as well as political and societal
constraints, the international community proposed and adopted numerous policy
initiatives that advocated switching from fossil fuels to energy produced by alterna-
tive energy sources (e.g., solar energy, hydrogen, and biofuels)gained more atten-
tion. Since 2006, both developed and developing nations have advanced and
continued to pursue pro-biofuel strategies and regulations (UNCTAD, 2023).
As of 2011, there are national mandates for mixing biofuels in 31 countries
across the globe and in 29 states and provinces (REN21, 2011). For example, both
the United States and the European Union approved legislation requiring significant
increases in biofuel consumption over the course of the next decade (Mitchell,
2011). Specifically, the US Energy Independence and Security Act of 2007 included
economic incentives and expanded the Renewable Fuel Standard to increase biofuel
production to 36 billion gallons by 2022, of which 21 billion gallons were required
to come from cellulosic biofuel or advanced biofuels derived from feedstocks other
than cornstarch (USEPA, 2023b). Additionally, to curb GHG emissions, the act
states that cellulosic biofuels must reduce GHG emissions by 60%, biodiesel and
advanced biofuels must reduce emissions by 50%, and conventional renewable
fuels (such as corn starch ethanol) must reduce life-cycle GHG emissions relative to
life-cycle emissions from fossil fuels by at least 20% (USEPA, 2023b). The 2009
European Energy and Climate Change Package set out a 10% minimum target for
renewable energy consumed by the transport sector to be achieved by all European
Union member states in their countries by the year 2020 (USDA, 2022). Additionally,
the Renewable Energy Sources Directive and Fuel Quality Directive specify the
minimum required amount of biofuel to be added to the motor fuel used in the
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 3

transport sector (Küüt et al., 2017). The Indian Ministry of Petroleum and Natural
Gas published its “National Policy on Biofuels” in 2018, and it was further updated
in 2022 (IEA, 2023). The goal of the policy is to increase domestic biofuel produc-
tion while reducing petroleum product imports. In China, in response to then-­
Premier Li Keqiang’s speech promoting ethanol use in 2018, numerous Chinese
provinces announced new or extended regulations to establish fuel gasoline-ethanol
blending schemes by 2020 (USDA, 2019).
Owing to concerns over climate change impacts resulting from fossil fuel com-
bustion, the use of biofuels has been proposed as an important solution to this recent
energy crisis (Rosillo-Calle, 2022). In general, biofuels are promoted as a low-­
carbon alternative to fossil fuels as they emit significantly lower GHG compared
with their fossil fuel-derived counterparts and, unlike other fuel additives, are fully
biodegradable (USDE, n.d.; OEERE, 2013). One key benefit of using biofuel is that
it can lower overall CO2 emissions without significantly altering our current infra-
structure (MIT, 2023). Some biofuels have significant physicochemical overlap
between the characteristics of a conventional petroleum-derived fuel, and they are
compatible with existing engine and fuel requirements (Pillay et al., 2008). In con-
trast to some petroleum-based products, they are not regarded as toxic or harmful;
hence, they are relatively simple to transport or store. Although biofuels have sev-
eral advantages over conventional fuels, there are still some potential drawbacks.
For example, on a per-unit energy basis, biofuels are often more expensive to pro-
duce than fossil fuels. Additionally, land and water resources available for food
production are in competition with feedstock growth to produce biofuel (Pachapur,
2020). In addition, the local climate, geographic location, soil fertility, and agricul-
tural practices of a region can impact the availability of biomass feedstock for pro-
ducing certain biofuels (USEPA, 2023b; Datta et al., 2019). Changes in land-use
patterns can lead to an increase in GHG emissions, stress on water supplies, pollu-
tion of the environmental reservoirs, and higher food prices. However, researchers
have argued that because biofuels are created from renewable feedstocks, as opposed
to fossil fuels, biofuel production and use could, in theory, be sustained indefinitely
(USEPA, 2023b).
In an effort to determine whether biomass energy is a boon or bane for society,
this review chapter focuses on a thorough and systematic review of previously pub-
lished material on biomass energy usage and its impact on the surrounding environ-
ment. In terms of the organization of this chapter, following the introduction, Sect.
2 focuses on methodology, and Sect. 3 presents a discussion on various biomass
feedstocks (including waste biomass) and the classification of biofuels. Next, Sect.
4 discusses climate change, GHGs, and biofuels. Subsequently, Sect. 5 presents the
findings of the literature review on biomass and biofuel. Section 6 presents biofuel
trends in the United States, India, and China, the world’s biggest GHG emitters.
Finally, Sect. 7, inspired by the United Nations Sustainable Development Goals
(SDGs) on the importance of ensuring access to clean and affordable energy (SDG
7) and climate change mitigation (SGD 13), examines whether biomass waste is a
blessing or a curse for society.
4 S. E. Haque and T. R. Rafi

2 Methodology

This chapter presents a systematic literature review of available scholarly and non-­
scholarly material on the topic. Specifically, the authors extracted 97 articles and
systematically (i) developed research questions to direct the study; (ii) searched the
most relevant in various databases (e.g., Science Direct, Web of Science, Scopus,
and relevant academic journals); (iii) assessed the selected articles’ quality and rel-
evance; (iv) summarized the scientific findings; and (v) analyzed/interpreted the
results.

3 Biomass Feedstock and Classification of Biofuels

Biomass feedstocks for energy production include agriculture products (including


algae/kelp) and waste, forestry residues, wood processing residue sorted municipal
solid wastes, sewage sludge, and animal waste (Fig. 1). Biomass can either be grown
for feedstock purposes or be a residue, such as wood waste from the logging indus-
try, which is usually left unused. Biomass wastes can be collected from agricultural
and forestry wastes, animal wastes, industrial wastes, and municipal solid wastes as
low-cost raw materials (Cho et al., 2020; Zhou & Wang, 2020). Note that of 4 bil-
lion hectares of forests worldwide, approximately half falls within developing

Fig. 1 Some common sources of biomass feedstocks


Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 5

nations (PBL, 2014). The generation of recovery of residue and processing waste
relies on factors such as tree species and local geographical conditions, and 1 m3 of
waste remains in the forest for every cubic meter of logged material removed.
Instead of fully relying on dedicated energy crops and agricultural crops, waste
biomass can also be utilized to sustainably meet the energy needs.
Biofuel can be in solid, liquid, and gaseous phases. Typically, nonfossil, organic
materials such as firewood, charcoal, and municipal trash, are used to make solid
biofuels (Manandhar et al., 2022). Biodiesel is a broad term used to describe a vari-
ety of liquid biofuels including methanol, ethanol, organic oils, and methyl esters
(Chiaramonti et al., 2007). Gaseous biofuels (e.g., methane) are primarily generated
from the pyrolysis or gasification of agricultural wastes and wood, as well as the
fermentation of animal manure (Datta et al., 2019; Szwaja et al., 2013).
Biofuels may be classified into two types: primary and secondary biofuels
(Table 1). Burning woody or cellulose plant material and animal manure directly
produce the primary biofuels, which utilize organic materials in their natural, unpro-
cessed state (Rodionova et al., 2017). Due to their unrefined nature, generally, the
primary biofuels are inefficient and pose a threat to the environment (Day et al.,
2014). The secondary biofuels are further divided intofour generations, according to
feedstock and/or biosynthetic platform (Alia et al., 2019; Rodionova et al., 2017).
The first generation of biofuels is ethanol, mostly produced from edible biomass
rich in starch, and biodiesel, primarily derived from vegetable oils and waste animal
fats such as cooking grease (Rodionova et al., 2017). The first generation of biofuels
has an adverse impact on food security because they are produced from edible plant
parts. Such limitations open the door for the production of second-generation biofu-
els, which are produced from various feedstocks, including nonfood lignocellulosic
biomasses (Raghavendra et al., 2019; Ullah et al., 2015). Additionally, second-­
generation biofuels have lower CO2 emissions than first-generation biofuels and
fossil fuels since the growth of cellulosic biomass does not require any additional
agrochemicals, water, or land-use changes (Fekete, 2013).Second-generation biofu-
els became a commercial reality since 2015 as countries made commitments toward
a more environmentally balanced future through the Sustainable Development
Goals (SDGs) and the Paris COP21 climate change agreement (UNCTAD, 2016).
The third generation of biofuels is produced from microorganisms, microalgae, and
seaweeds, and it is a promising way of meeting global demands (Alia et al., 2019;
Rodionova et al., 2017). The latest biofuel generation, the fourth-generation, feed-
stock includes genetically modified organisms (e.g., cyanobacteria) for the produc-
tion of ethanol along with fuel products such as butanol, isobutanol, and modified
fatty acids (Goria et al., 2022; Dexter & Fu, 2009). For example, Cyanobacteria are
modified to boost oil yield and for efficient bioenergy production (Simya et al.,
2018). Additionally, the feedstocks can be grown on non-arable land, and the appli-
cation of bioengineering concepts to alter the characteristics and metabolism of
algae increases the oil quantity in the cells. Moreover, the higher oil yield aids in
atmospheric CO2 reduction.
Table 1 Classification of biofuels derived from various biomass feedstocks (Suali & Suali, 2023; USEPA, 2023b; Goria et al., 2022; Ofori-Boateng, 2022; Ziolkowska,
2020; Alia et al., 2019; Datta et al., 2019; Kumar et al., 2018; Rodionova et al., 2017; Aro, 2015; Day et al., 2014; S&TCI, 2006)
Secondary biofuels
Third generation Fourth generation of
Primary biofuels First generation of biofuels Second generation of biofuels of biofuels biofuels
Produced Burning woody or Sugar crops, starch crops, oilseed crops, and Nonfood crops including the waste from Microorganisms, Genetically modified
from cellulose plant animal fats food crops, agricultural residue, wood microalgae, and organisms
material, dry animal chips, and waste cooking oil seed seaweed
manure
Products Substituted for Corn ethanol, biodiesel, and pure plant oils Jatropha, mahua, pongamia, neem, rubber, Biodiesel, Ethanol, butanol,
conventional fossil karanja, castor oil, etc. gasoline, butanol, isobutanol, and
fuel in heating, propanol, and modified fatty acids
cooking, or ethanol
electricity
production
Remarks (i) Owing to their (i) Relatively low unit production investment (i) An improved method, which is (i) Regarded as (i) Aims to provide
unrefined nature, requirements, and simple and well-known developed to overcome the constraints superior more sustainable
in general, these processing technologies of original biofuels biofuel production options
biofuels are (ii) Adverse impact on food supply, food (ii) As the inevitable by-product of the substitutes as by combining
inefficient and security, and arable land requirements agricultural industry is used as raw these biofuels biofuels production
create a negative (iii) Limited profit margin compared to fossil material, no further agrochemical water can almost with the capture and
impact on fuels in terms of GHGs as they require a or land is needed to grow the feedstock entirely avoid storage CO2
environment significant amount of energy to produce, (iii) The majority of the second-­generation the drawbacks (ii) Involves using
store, and utilize biofuels and the ability to utilize less of first- and genetic engineering
(iv) The majority of automobiles can run on expensive feedstocks do not second-­ to increase the
gasoline-­ethanol mixtures with up to necessarily result in lower-cost generation preferred
10% ethanol (by volume). E85 is a biofuels. The lower feedstock costs are biofuels characteristics of
gasoline-­ethanol mixture that contains up countered by increased chemical costs organisms used in
to 85% ethanol and is suitable for use in and significantly increased estimated biofuel production
flexible fuel cars capital expenses
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 7

4 Climate Change, Greenhouse Gases, and Biofuels

Burek (2010) assessed historical data from 1985 on consumption trends, and the
known fossil fuel supplies found that, based on the trends at the time of the study,
all fossil fuels could run out within decades, possibly as early as 2060. The predic-
tion is based on both the increase in energy usage and the reserves of fuels.
Additionally, a recent report published by the International Energy Agency (IEA,
2023) indicates that between 2023 and 2025, the world’s electricity demand is antic-
ipated to increase at a rate of 3% annually, which is significantly higher than the
growth rate of 2022. In 2022, the effect of China’s zero-COVID policy on its econ-
omy was significant, and uncertainty remains regarding the country’s electricity
demand. However, while COVID-19-related restrictions impacted China’s growth,
the demand for electricity increased in both India and the United States (IEA, 2023).
In an effort to minimize and prevent any community spread of the coronavirus,
China carried out widespread testing, housed the effected individuals in government
facilities, and established targeted lockdowns. Whereas during the same year in
India, a combination of strong post-pandemic economic rebound and high summer
temperatures led to an 8.4% increase in electricity demand (IEA, 2023). Furthermore,
during 2022, energy demand in the United States increased substantially by 2.6%
year over year, driven by economic activity and higher residential consumption to
meet heating and cooling needs during hotter summers and colder-than-average
winters. Isaac and van Vuuren (2009) studied the residential sector energy demand
worldwide for heating and air conditioning considering the impacts of changing
climate. The findings of this study indicate that due to the impacts of global climate
change, by 2100, there will be a 34% decrease in the demand for heating and a 72%
increase in the demand for air cooling. van Ruijven et al. (2019) investigated the
growth in future energy consumption based on an analysis that considered Gross
Domestic Product and temperature and 210 socioeconomic and climate scenarios.
These researchers found that in the future, energy consumption will possibly
increase owing to the impacts of global climate change; however, the magnitude is
dependent on several interacting sources of uncertainty.
Currently, the primary source of atmospheric CO2 emissions comes from the
power sector due to fossil fuel combustion, and by far, fossil fuels are the dominant
cause of climate change, producing over 75% of all GHG emissions and almost
90% of all CO2 emissions (UN, n.d). It is now well established that due to the
impacts of climate change and global warming, the earth’s surface temperature is
increasing (Haque, 2023; Haque & Nahar, 2023; IPCC, 2007). Since 1880, our
planet’s temperature has risen by approximately 1.1 °C (NASA, 2022) and the
growth rate has more than doubled over the last 40 years (0.18 °C per decade;
Lindsey & Dahlman, 2021). The Intergovernmental Panel on Climate Change
(IPCC, 2007) predicts future warming trends, and by 2100, it is predicted that global
average surface temperatures would rise by 1.1–6.4 °C; however, we can change the
trajectory of global temperature increases by taking immediate action (IPCC, 2022).
8 S. E. Haque and T. R. Rafi

For example, the utilization of biomass, as a source of renewable energy, and a tra-
ditional energy source can be vital in minimizing the environmental effect of fossil
fuel combustion to produce energy (McKendry, 2002). However, it is noteworthy
thatrecently, the USEPA (2023b) reported that depending on the feedstock, the pro-
duction process, and the time horizon of the analysis, biofuels can emit even more
GHGs per unit of energy than some fossil fuels. Additionally, although biofuels
have environmental benefits, their production and use can also have adverse impact
on the environment. For example, if spilled, pure ethanol and biodiesel break down
into harmless substances, but fuel ethanol, which contains denaturants to make fuel
ethanol undrinkable, is flammable (particularly ethanol; EIA, 2022).

5 Summary of Previous Research Findings


on Biomass Energy

In recent years, there has been a huge quantity of research in the area of biomass and
biofuel. Table 2 presents a summary of the previously published research works
related to biomass and biofuel.

6 Biofuels in the United States, India, and China

The findings of this review indicate that globally, there are numerous biomass
sources available to produce biofuels, and the biomass feedstock used is country-­
specific. Of the 4 billion hectares of forest worldwide, approximately half falls
within developing countries. Since the early 1980s, the production and consumption
of biofuels have generally increased annually in the United States, and various gov-
ernment policies and programs encouraged and/or mandated the use of biofuels
(EIA, 2022). These policies and programs primarily aim to reduce the usage of fos-
sil fuel-based transportation fuels. In the mid-2000s, following the initial phase of
explosive growth, biodiesel production capacity kept growing in the United States,
but in recent years, the rapid expansion of renewable diesel production capacity has
threatened its potential (Gerveni et al., 2023). This surge in renewable diesel pro-
duction raised a number of questions regarding the effect on biofuel, grain, and
oilseed markets (Gerveni et al., 2023). On the other hand, China and India stand out
with their high population density and notable food vs. fuel debates. The importance
of biofuels in replacing conventional transportation fuels in these two nations has
recently been highlighted in numerous studies; however, most of the work focuses
on unconventional processes such as lignocellulosic feedstocks (Beckman et al.,
2018). The current status and challenges for biofuels in the United States, India, and
China are briefly described in the sections below.
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 9

Table 2 Summary of previous research findings


Authors Year Summary of findings
Cavelius et al. 2023 This study investigates the potential of biofuels from first to fourth
generation. The findings reveal that it is necessary to develop renewable
energy sources, of which biofuels will contribute significantly.
Additionally, the use of biofuels will contribute to curbing CO2
emissions and meet the ever-increasing demand for energy. Further, the
findings suggest that to reduce the effects of climate change and make
the transition to a sustainable society, people are prepared to accept
change from the status quo
Osman et al. 2021 This review provides a critical assessment of current biomass to biofuel
conversion pathways and related studies, which investigate
environmental effects throughout the life cycle. The finding indicates
that the research on biochemical conversion is significantly less
explored, outweighed, and understood compared to the thermochemical
conversion of biomass
Shokravi et al. 2022 This study focuses on fourth-generation biofuel from genetically
modified algal biomass. According to the findings, the technical and
biosafety aspects of fourth-generation biofuel, along with the
complexity and diversity of the pertinent regulations, legal concerns,
and health and environmental impacts, are among the most pressing
issues, which call for a strong commitment at the national and
international levels to reach an agreement
Subramaniam 2021 The study aims to investigate the impact of economic globalization on
and Masron biofuel production using panel data from 50 developing nations between
2012 and 2016. The findings revealed that economic globalization has a
favorable effect on the production of biofuels, and the evidence is robust
to several robustness checks. According to the authors, promoting
globalization’s economic features will promote biofuel usage and reduce
their adverse influence on the environment
Saravankumar 2019 This study investigates the possibility of adding silicon dioxide
et al. nanoparticles to maize oil methyl ester in the form of an emulsion.
According to experimental findings, adding nanoparticles improves
emission characteristics in diesel engines as they serve as an oxidation
catalyst
Uddin et al. 2019 These researchers examined the potential of biomass energy as a source
of sustainable energy in Bangladesh. Additionally, the study aims to
predict the challenges associated with the future implementation of
biomass energy in the country. Furthermore, due to its wide availability,
biomass can be used as a sustainable energy source with support from
various levels of the government
Oliveira et al. 2017 These researchers investigate the political and financial motivations
behind biofuel policy and their influence on government interventions
and policy development. The findings reveal that biofuel policies and
processes are ineffective because they are not created with the interests
of the environment or pro-poor development in mind. Instead, they are
established and put into effect to serve the goals of corporations seeking
to maximize profits and governments worried about energy security
(continued)
10 S. E. Haque and T. R. Rafi

Table 2 (continued)
Authors Year Summary of findings
Rago et al. 2018 This study provides a review of thermochemical technologies for the
transformation of biomass waste to biofuel and energy in developing
nations. Economic, technological, and societal concerns account for the
developing nations’ resistance to the use of biomass thermochemical
conversion methods. In the long run, the thermochemical conversion of
biomass to biofuel can become a significant Clean Development
Mechanism project with curbed GHG emissions
Ji and Long 2016 This study offers a thorough and current assessment of the literature on
the socioeconomic and ecological impacts of biofuels. The research
reveals that the conclusions of published literature are inconsistent
regarding the ecological or socioeconomic effects of biofuels, and there
is doubt about the cleanliness and renewability of these fuels
Sekoai and 2016 These researchers studied the development initiatives of biofuels in
Yoro sub-Saharan Africa. The findings indicate that biofuel development
initiatives have been stagnant in Africa owing to governmental
regulations, and a lack of funding, technological know-how, and
available land. Additionally, the results imply that biofuel projects will
act as a stimulant for Africa’s economic expansion, infrastructure
improvement, and social welfare
Su et al. 2015 These researchers review the national biofuel policies and strategy plans
of the leading nations across the globe. The results reveal that although
there is political disagreement regarding the effect of biofuels on food
security and global climate change, it would be difficult for
policymakers to maintain trends of biofuel production while adhering to
rules for sustainable production

6.1 United States

Recently, the United States passed legislation and set ambitious goals to promote
the development and commercialization of biofuels (USEPA, 2023b; Hoekman,
2009). Additionally, with regard to biofuels, individual US states have adopted
aggressive approaches, with initial attempts primarily concentrated on ethanol, cre-
ated through the fermentation of carbohydrates from cereals, particularly corn
(Hoekman, 2009). Moreover, the US Energy Independence and Security Act pro-
vided different subsidies, donations, loans, and cash awards to assist biofuel research
and gave incentives for biorefineries that helped to replace 80% of fossil fuels and
use cellulosic ethanol to assure the sustainable use of biofuel in the US economy
(Sajid et al., 2021). According to the US Energy Information Administration (2022),
in 2021, the country produced approximately 17.5 billion gallons of biofuels and
consumed roughly 16.8 billion gallons. During the same year, the country exported
0.8 billion gallons of biofuels on a net basis, with gasoline-ethanol making up the
lion’s share of both gross and net exports. Furthermore, the majority of biofuel is
consumed in blends with refined petroleum products including gasoline, diesel fuel,
heating oil, and kerosene-type jet fuel.
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 11

It is important to note that for a developed country such as the United States, the
benefits of biofuel include energy security, economic growth, and environmental
protection; nonetheless, there are a number of challenges, which must be overcome
before these advantages can be realized. For instance, DellaSalla and Koopman
(2016) investigated biomass energy generation and forest thinning in the western
United States and its influence on GHG emissions. These researchers found that in
the western portion of the United States, biomass energy created as a by-product of
forest clearing is being promoted as a “win-win” for lowering fire risks and substi-
tuting fossil fuels. However, many justifications regarding thinning and biomass
approaches require validation regarding whether they are in fact ecologically sound
and carbon neutral. In addition, extensive forest thinning and energy production
from forest biomass without adequate safeguards are highly risky strategies for lim-
iting the influences of changing climate with potentially irreversible consequences
to fire-adapted forests and GHG emissions.

6.2 India

Usmani (2020) studied the potential to generate biofuel from biomass in India,
South Asia’s largest economy. India has a population of over 1.2 billion, and for the
next several decades, the country’s population is projected to grow at an unprece-
dented rate (Usmani, 2020). With a demand of 5 million barrels/day, India is the
third-largest consumer behind the United States and China (HT, 2022). In particular,
the Indian transport industry consumes 99.6% of petrol and approximately 70% of
diesel (Usmani, 2020). The total quantities of petroleum consumed in India in
2019–20 were 194.3 million metric tons (MMT), up around 5% to 204.2 MMT (HT,
2022). As of 2018, imports accounted for more than 80% of the nation’s oil needs
(TET, 2018), indicating that the country’s domestic production of crude oil could
only satisfy 20% of the country’s needs. Additionally, India is projected to exhaust
its coal reserve and its primary source of energy, over the next three decades (Patni
et al., 2011). Its domestic natural gas reserves are limited as well. Numerous
researchers indicate that the country’s fuel energy security would continue to be at
risk unless alternative fuels are produced based on renewable feedstocks (Patni
et al., 2011).
The Indian Ministry of Petroleum & Natural Gas (MPNG, 2023) considers bio-
fuels as the solution to insufficient fossil fuel supply and GHG emission problems.
According to a 2011 Asian Development Bank (ADB, 2017) study, first-generation
bioethanol has limited market potential in the country because it competes with
agricultural resources and threatens food security. The study further indicates that a
mix of policies, such as expanding biodiesel production, increasing energy effi-
ciency, and increasing food productivity, will give India better food and energy
security and chances for inclusive growth and carbon emission reduction. By 2030,
the Indian government plans to reduce the nation’s carbon footprint by 30–35%
12 S. E. Haque and T. R. Rafi

(TET, 2020). This goal will be attained through a five-pronged plan, which envi-
sions biofuels playing a strategic role in India’s energy mix. In addition, by 2030,
the Indian Government envisages an indicative target of blending 20% ethanol and
5% biodiesel into gasoline and diesel, respectively (MPNG, 2023). Moreover, the
government has announced several efforts to boost domestic biofuel production.
Liquid biofuels may be used in combination with petroleum fuels without modify-
ing engines, which makes their widespread implementation quite simple (Usmani,
2020). Furthermore, Karunakaran (2023) reports that biofuel made from agricul-
tural waste offers India’s farmers the great potential to be empowered, revolution-
izes the rural economy, and enhances the quality of life in rural regions by creating
revenue and generating employment opportunities. However, India has restrictions
on the types of feedstocks that can be used to make biofuels, such as banning the use
of sugarcane juice to make ethanol, which has led to the slow expansion of biofuel
production in India (Beckman et al., 2018).

6.3 China

China, the largest developing nation in the world, is the world’s top energy con-
sumer (Chen et al., 2020). Kang et al. (2020) investigated bioenergy in China and
found that domestic biomass resources’ collectible potential increased from approx-
imately 18.3 exajoule (1018 J; EJ) in 2000 to 22.7 EJ in 2016. In 2016, the entire
potential for energy crops (32.7EJ) amounted to roughly 27.6% of China’s energy
consumption. The cumulative reduction in GHG emissions due to this potential, if
it can be realized strategically to replace fossil fuels between 2020 and 2050, would
be in the range of about 652.7–5859.6 Mt CO2-equivalent, with the negative GHG
emissions attributable to the introduction of bioenergy with carbon capture and stor-
age accounting for 923.8–1344.1 Mt CO2-equivalent. Due to its limited energy
resources and excessive reliance on burning fossil fuels for energy generation,
China is concerned about its energy security, which puts pressure on the Chinese
government to adjust its energy mix (Zhang et al., 2020). The government is pro-
moting the national biofuel initiative to address issues with energy security; how-
ever, grain-based ethanol is no longer supported by Chinese policy, which prohibits
biofuel feedstocks from competing with feedstocks for human or animal use
(Beckman et al., 2018; Koizumi, 2013). Moreover, Chen et al. (2020) studied micro-
algal biofuels in China and reported that China is working toward creating a variety
of renewable energy sources and has made substantial investments in microalgal
biomass and biofuel production. Additionally, to address the country’s energy and
environmental concerns, China has initiated several significant research and devel-
opment programs for microalgal biomass and biofuels. Weng et al. (2019) report
that planting energy crops on marginal land might increase non-grain feedstocks by
10% and save approximately 0.22% of croplands, reducing the adverse influences
on land resources and food security (Weng et al., 2019). As a result, this may be one
of the promising routes for sustainable biofuel development in the country.
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 13

7 Biofuels: A Boon or Bane for Society?

The declining market share of natural fossil fuels necessitated research organiza-
tions, decision-makers, and businesses to find other ways to provide transportation
fuel. Biomass energy is a promising solution to this problem, which some research-
ers suggest assists in climate change mitigation while producing significant quanti-
ties of energy through a relatively simpler process (Surriya et al., 2015). Past
researchers found that better land management, job creation, the use of underuti-
lized agricultural land in industrialized nations, the utilization of modern energy
sources to rural communities in developing nations, lowing of atmospheric CO2
concentrations, improved waste management, and the recycling of nutrients are
some of the benefits of biomass energy (Zhang et al., 2020). Even though biofuels
appear to be a sustainable solution for ensuring energy supplies and an eco-friendly
solution for curbing atmospheric CO2 emissions, there are serious concerns regard-
ing their growing use (Jeswani et al., 2020; Azapagic, 2011). For example, biodiver-
sity, an environmental component of sustainability, can be impacted by land
conversion to produce enough biomass to significantly reduce fossil fuel depen-
dency (Araújo et al., 2017). Silvestri (2008) conducted research on Honduran bio-
diesel generation using palm oil. The results of this study revealed that the production
of palm oil-based biodiesel did not present new employment opportunities for
Honduran rural poor communities and was more likely to affect food security while
adversely influencing the environment. In addition, the findings indicate that bio-
diesel production from palm oil is not as advantageous as it might seem. A separate
study found that the production of feedstocks and yields have both gradually but
steadily increased over the last several years, along with the share of fuels based on
bioenergy (Ajanovic, 2011).
The current policy question revolves around whether the trade-offs between food
and fuel are less harmful or unwarranted for developing countries and which aspects
are likely to be critical in determining future biofuel endorsements (Das &
Gundimeda, 2022). It would be beneficial to describe the many generations of bio-
fuels that have entered the market before examining the sustainability of particular
biofuels. Jeswani et al. (2020) reviewed the environmental sustainability of biofuels
and found that the use of first-generation feedstocks has become an especially sensi-
tive topic because of the competition with food production and concerns regarding
using agricultural land for biofuel production. Growing agricultural product demand
puts the environment in danger by increasing the likelihood of deforestation, the
need to use land with high biodiversity values to meet this demand, and the usage of
freshwater and agrochemicals that go along with it. Additionally, increasing food
and fuel prices are anticipated to endanger the food security of nations, which
import both food and gasoline, whereas in countries, which are net exporters in one
and net importers of the other, the situation will be governed by the relative size of
the food or energy exports and imports (USDA, 2007). Further, Lackner (2022)
indicates that owing to land-use changes, fertilizer consumption, and process yields,
first-generation biofuels can perform even worse than petroleum-based fuels in
14 S. E. Haque and T. R. Rafi

terms of their net influence on changing climate. Although some of these concerns
could be resolved by employing second-generation feedstocks, the practicality of
some second-generation biofuels is still debatable, primarily because they are sus-
tainably more expensive than petroleum fuels on an energy equivalent basis
(Carriquiry et al., 2011). The third generation of biofuels can bypass the issues of
food competition and land utilization as microalgae show great promise for future
biofuel generation in the saline, brackish, and freshwater environments as they may
form up to an order of magnitude more biomass per area than terrestrial biomass
(Lackner, 2022; Jeswani et al., 2020). However, biofuel production from microalgae
is energy-intensive and currently not profitable (Passell et al., 2013).
Overall, the findings of the review indicate that while biofuels primarily help to
attain SDGs 7 (energy supply security and reduction in fossil fuel use) and 13 (ame-
liorate global warming due to reduction in atmospheric CO2 emission), they also
have detrimental effects on other such as SDG 2 (Zero hunger) and SDG 12 (respon-
sible consumption and production), which vary according to use of type of biomass
feedstock to produced biofuels (Nazari et al., 2020). In the end, the review finds that
the issue of biofuel usage is a complex interplay of two different worldviews; on the
one hand, the developed nations consider biofuels as renewable and eco-friendly,
whereas the developing nations are focused on having food on the table than pro-
tecting the environment.
The findings of this review reveal that further research and development are
required before determining whether biofuels are a boon or a bane for society.

Acknowledgments The authors are appreciative of their families for supporting them during the
many hours of work spent writing this chapter.

References

ADB. (2017). Food security, energy security, and inclusive growth in India: The role of biofuels.
Available at: https://www.adb.org/publications/food-­security-­energy-­security-­and-­inclusive-­
growth-­india-­role-­biofuels. Accessed: July 8, 2023.
Ajanovic, A. (2011). Biofuels versus food production: Does biofuels production increase food
prices? Energy, 36(4), 2070–2076. Available at: https://doi.org/10.1016/j.energy.2010.05.019.
Accessed: June 9, 2023.
Alia, K. B., Rasul, I., Azeem, F., et al. (2019). Microbial production of ethanol. Microbial fuel
cells: Materials and applications. Materials Research Foundations, 46, 307–334. Available at:
https://doi.org/10.21741/9781644900116-­12. Accessed: June 9, 2023.
Araújo, K., Mahajan, D., Kerr, R., & da Silva, M. (2017). Global biofuels at the crossroads:
An overview of technical, policy, and investment complexities in the sustainability of bio-
fuel development. Agriculture, 7. Available at: https://doi.org/10.3390/agriculture7040032.
Accessed: June 16, 2023.
Aro, E.-M. (2015). From first generation biofuels to advanced solar biofuels. Ambio: A Journal of the
Human Environment, 45(S1), 24–31. Available at: https://doi.org/10.1007/s13280-­015-­0730-­0.
Accessed: June 16, 2023.
Azapagic, A. (2011). Assessing the sustainability of biofuels. Available at: https://research.man-
chester.ac.uk/en/publications/assessing-­sustainability-­of-­biofuels. Accessed: July 3, 2023.
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 15

Barbieri, L., Andreola, F., Lancellotti, I., & Taurino, R. (2013). Management of agricultural bio-
mass wastes: Preliminary study on characterization and valorisation in clay matrix bricks. Waste
Management, 33(11), 2307–2315. Available at: https://doi.org/10.1016/j.wasman.2013.03.014.
Accessed: June 28, 2023.
Beckman, J., Gooch, E., Gopinath, M., & Landes, M. (2018). Market impacts of China and India
meeting biofuel targets using traditional feedstocks. Biomass & Bioenergy, 108, 258–264.
Available at: https://doi.org/10.1016/j.biombioe.2017.11.018. Accessed: June 28, 2023.
Burek, S. (2010). When will fossil fuels finally run out and what is the technical potential for
renewable energy resources? International Journal of COMADEM, 13(4), 22–27.
Carriquiry, M. A., Du, X., & Timilsina, G. R. (2011). Second generation biofuels: Economics and
policies. Energy Policy, 39(7), 4222–4234. https://doi.org/10.1016/j.enpol.2011.04.036
Cavelius, P., Englehart-Straub, S., Mehlmer, N., Lercher, J., Awaf, D., & Brück, T. (2023).
The potential of biofuels from first to fourth generation. PLoS Biology, 21(3). https://doi.
org/10.1371/journal.pbio.3002063
Chen, H. S., Wang, X., & Wang, Q. (2020). Microalgal biofuels in China: The past, progress
and prospects. GCB Bioenergy, 12(12), 1044–1065. Available at: https://doi.org/10.1111/
gcbb.12741
Chiaramonti, D., Oasmaa, A., & Solantausta, Y. (2007). Power generation using fast pyrolysis
liquids from biomass. Renewable & Sustainable Energy Reviews, 11(6), 1056–1086. Available
at: https://doi.org/10.1016/j.rser.2005.07.008. Accessed: June 27, 2023.
Cho, E. J., Trinh, L. T. P., Song, Y., et al. (2020). Bioconversion of biomass waste into high value
chemicals. Bioresource Technology, 298, 122386.
Das, P., & Gundimeda, H. (2022). Is biofuel expansion in developing countries reasonable? A
review of empirical evidence of food and land use impacts. Journal of Cleaner Production,
372. Available at: https://doi.org/10.1016/j.jclepro.2022.133501. Accessed: June 4, 2023.
Datta, A., Hossain, A., & Roy, S. D. (2019). An overview on biofuels and their advantages and
disadvantages. Asian Journal of Chemistry, 31(8), 1851–1858. Available at: https://doi.
org/10.14233/ajchem.2019.22098
Day, C., Tseng, Y.-C., Puyol, R., & Nissan, J. (2014). Efficiency comparisons of secondary biofu-
els. PAM Review, 1, 70–89. Available at: https://doi.org/10.5130/pamr.v1i0.1386
DellaSalla, D. A., & Koopman, M. (2016). Thinning combined with biomass energy produc-
tion impacts fire-adapted forests in western United States and may increase greenhouse gas
­emissions. In Reference module in earth systems and environmental sciences. Elsevier. https://
doi.org/10.1016/B978-­0-­12-­409548-­9.09587-­7
Dexter, J., & Fu, P. (2009). Metabolic engineering of cyanobacteria for ethanol production. Energy
and Environmental Science, 2(8), 857. https://doi.org/10.1039/b811937f
EIA. (2022). Biomass explained. Available at: https://www.eia.gov/energyexplained/
biomass/#:~:text=Direct%20combustion%20is%20the%20most,biomass%20includes%20
pyrolysis%20and%20gasification. Accessed: June 25, 2023.
Fekete, B. M. (2013). Biomass. In Elsevier eBooks (pp. 83–87). s.n.
Gerveni, M., Hubbs, T., & Irwin, S. (2023). Overview of the production capacity of U.S. bio-
diesel plants. farmdoc daily, (13), 32. Department of Agricultural and Consumer Economics,
University of Illinois at Urbana-Champaign, February 22, 2023.
Goria, K., Kothari, R., Singh, H. M., Singh, A., & Tyagi, V. V. (2022). Biohydrogen: Potential
applications, approaches, and hurdles to overcome. In Elsevier eBooks (pp. 399–418). Available
at: https://doi.org/10.1016/b978-­0-­12-­822810-­4.00020-­8
Guan, Y., Yan, J., Shan, Y., Zhou, Y., Hang, Y., et al. (2023). Burden of the global energy price crisis
on households. Nature Energy, 8(3), 304–316. https://doi.org/10.1038/s41560-­023-­01209-­8
Haque, S. E. (2023). Chapter 2 – Historical perspectives on climate change and its influence on
nature. In A. Srivastav, A. Dubey, A. Kumar, S. K. Narang, & M. A. Khan (Eds.), Visualization
techniques for climate change with machine learning and artificial intelligence (pp. 15–38).
Elsevier. https://doi.org/10.1016/B978-­0-­323-­99714-­0.00003-­0
16 S. E. Haque and T. R. Rafi

Haque, S. E., & Nahar, N. (2023). Bangladesh: Climate change issues, mitigation, and adaptation in
the water sector, ACS EST Water, 3, 6, 1484–1501. https://doi.org/10.1021/acsestwater.2c00450
Hoekman, S. K. (2009). Biofuels in the U.S. – Challenges and opportunities. Renewable Energy,
34(1), 14–22. Available at: https://doi.org/10.1016/j.renene.2008.04.030
HT (Hindustan Times). (2022). India’s daily petroleum consumption growing faster than
global average: Puri. https://www.hindustantimes.com/business/indias-­daily-­petroleum-­
consumption-­growing-­faster-­than-­global-­average-­puri-­101665762596340.html. Accessed on
July 17, 2023.
IEA. (2011). https://www.iea.org/news/biofuels-­can-­provide-­up-­to-­27-­of-­world-­transportation-­
fuel-­by-­2050-­iea-­report-­says-­iea-­roadmap-­shows-­how-­biofuel-­production-­can-­be-­expanded-­
in-­a-­sustainable-­way-­and-­identifies-­needed-­technologies-­and-­policy-­actions. Accessed on
July 16, 2023.
IEA. (2022). Global energy review: CO2 emissions in 2021. Available at: https://www.iea.org/
reports/global-­energy-­review-­co2-­emissions-­in-­2021-­2. Accessed: June 7, 2023.
IEA. (2023). National policy on biofuels (2022 Amendment). Available at: https://www.iea.org/
policies/17006-­national-­policy-­on-­biofuels-­2022-­amendment. Accessed: July 7, 2023.
IPCC. (2007). Climate change 2007: The physical science basis. Contribution of Working Group I
to the fourth assessment report of the Intergovernmental Panel on Climate Change. Cambridge
University Press. Accessed on June 12, 2023.
IPCC. (2022). Mitigation of climate change. https://www.ipcc.ch/report/sixthassessment-­report-­
working-­group-­3/. Accessed on June 17, 2023.
Isaac, M., & van Vuuren, D. P. (2009). Modeling global residential sector energy demand for
heating and air conditioning in the context of climate change. Energy Policy, 37(2), 507–521.
https://doi.org/10.1016/j.enpol.2008.09.051
Jeswani, H. K., Chilvers, A., & Azapagic, A. (2020). Environmental sustainability of biofuels: A
review. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences,
476(2243). Available at: https://doi.org/10.1098/rspa.2020.0351. Accessed: June 16, 2023.
Ji, X., & Long, X. (2016). A review of the ecological and socioeconomic effects of biofuel and
energy policy recommendations. Renewable and Sustainable Energy Reviews, 61, 41–52.
https://doi.org/10.1016/j.rser.2016.03.026
Jiang, P., Walmsley, T. G., & Klemeš, J. J. (2021). Impacts of COVID-19 on energy demand and
consumption: Challenges, lessons and emerging opportunities. Applied Energy, 285. Available
at: https://doi.org/10.1016/j.apenergy.2021.116441. Accessed: June 15, 2023.
Kang, Y., Yang, Q., Bartocci, P., Wei, H., Liu, S. S., et al. (2020). Bioenergy in China:
Evaluation of domestic biomass resources and the associated greenhouse gas mitiga-
tion potentials. Renewable and Sustainable Energy Reviews, 127, 109842. https://doi.
org/10.1016/j.rser.2020.109842
Karunakaran, K. (2023, January 18). Achieving India’s energy security and economic resilience
with biofuels. Times of India Blog. Available at: https://timesofindia.indiatimes.com/blogs/
voices/achieving-­indias-­energy-­security-­and-­economic-­resilience-­with-­biofuels/. Accessed:
June 20, 2023.
Koizumi, T. (2013). Biofuel and food security in China and Japan. Renewable & Sustainable
Energy Reviews, 21, 102–109. Available at: https://doi.org/10.1016/j.rser.2012.12.047.
Accessed: June 8, 2023.
Kumar, M., Babu, A. V., & Kumar, P. S. (2018). The impacts on combustion, performance and emis-
sions of biodiesel by using additives in direct injection diesel engine. Alexandria Engineering
Journal, 57(1), 509–516. Available at: https://doi.org/10.1016/j.aej.2016.12.016
Küüt, A., Ilves, R., Küüt, K., Raide, V., Ritslaid, K. M., & Oltm, J. (2017). Influence of European
Union directives on the use of liquid biofuel in the transport sector. 10th international scientific
conference Transbaltica 2017: Transportation science and technology. Procedia Engineering,
187, 30–39. https://doi.org/10.1016/j.proeng.2017.04.346
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 17

Lackner, M. (2022). Third-generation biofuels: Bacteria and algae for better yield
and sustainability. In Springer eBooks (pp. 1947–1986). Available at: https://doi.
org/10.1007/978-­3-­030-­72579-­2_90
Lindsey, R., Dahlman, L. (2021), Climate Change: Global Temperature. NOAA Climate.gov.
https://www.climate.gov/news-features/understanding-climate/climate-change-global-temper-
ature. Accessed January 24, 2024.
Manandhar, A., Mousavi-Avval, S. H., Tatum, J., Shrestha, E., Nazemi, P., et al. (2022). Solid
biofuels. In Elsevier eBooks (pp. 343–370). Available at: https://doi.org/10.1016/b978-­0-­12-
­819242-­9.00017-­8
McKendry, P. (2002). Energy production from biomass (part 1): Overview of biomass. Bioresource
Technology, 83(1), 37–46. Available at: https://doi.org/10.1016/s0960-­8524(01)00118-­3
Mitchell, D. (2011). Biofuels in Africa opportunities, prospects, and challenges. https://docu-
ments1.worldbank.org/curated/en/617361468201837240/pdf/584380PUB0ID181Afr
ica09780821385166.pdf. Accessed January 24, 2024.
MIT Climate Portal. Available at: https://climate.mit.edu/explainers/biofuel. Accessed: July
4, 2023.
MPNG. (2023). About biofuel. Available at: https://mopng.gov.in/en/refining/about-­bio-­fuel.
Accessed: July 3, 2023.
NASA. (2022). World of change: Global temperatures. Available at: https://earthobservatory.nasa.
gov/world-­of-­change/global-­temperatures. Accessed: June 5, 2023.
Nazari, M. T., Nazari, M. T., Mazutti, J., Basso, L. G., & Branli, L. (2020). Biofuels and their
connections with the sustainable development goals: A bibliometric and systematic review.
Environment, Development and Sustainability, 23(8), 11139–11156. Available at: https://doi.
org/10.1007/s10668-­020-­01110-­4
OEERE (Office of Energy Efficiency & Renewable Energy). (2013). Ethanol vs. Petroleum-based
fuel carbon emissions. https://www.energy.gov/eere/bioenergy/articles/ethanol-­vs-­petroleum-­
based-­fuel-­carbon-­emissions. Accessed on July 16, 2023.
Ofori-Boateng, C. (2022). Global profile and market potentials of the third-generation biofuels.
In Woodhead Publishing series in energy, 3rd generation biofuels (pp. 745–756). Woodhead
Publishing. https://doi.org/10.1016/B978-­0-­323-­90971-­6.00018-­8
Oliveira, G. L. T., McKay, B., & Plank, C. (2017). How biofuel policies backfire: Misguided goals,
inefficient mechanisms, and political-ecological blind spots. Energy Policy, 108, 765–775.
https://doi.org/10.1016/j.enpol.2017.03.036
Osman, A., Mehta, N., Elgarahy, A. M., Al-Hinai, A., Ml-Muhtaseb, A., et al. (2021). Conversion
of biomass to biofuels and life cycle assessment: A review. Environmental Chemistry Letters,
19(6), 4075–4118. Available at: https://doi.org/10.1007/s10311-­021-­01273-­0
Pachapur, P. (2020). Food security and sustainability. Available at: https://www.semanticscholar.
org/paper/Food-­S ecurity-­a nd-­S ustainability-­Pachapur-­Pachapur/ed034f35604381016e7
2f784290d968bbd00d2c1. Accessed: June 19, 2023.
Passell, H. D., Dhaliwal, H., Reno, M., Wu, B., Amotz, A. B., et al. (2013). Algae biodiesel life
cycle assessment using current commercial data. Journal of Environmental Management, 129,
103–111. Available at: https://doi.org/10.1016/j.jenvman.2013.06.055
Patni, N., Pillai, S. G., & Dwivedi, A. H. (2011). Analysis of current scenario of biofuels in India
specifically biodiesel and bio-ethanol. In Conference: International conference on current
trends in technology, NUiCONE 11.
PBL. (2014). Netherlands Environmental Assessment Agency Report. Integrated analysis
of global biomass flows in search of the sustainable potential for bioenergy production.
PBL Publication no. 1509. http://www.pbl.nl/sites/default/files/cms/publicaties/pbl-­2014-­
integrated-­a nalysis-­o f-­g lobal-­b iomass-­f lows-­i n-­s earch-­o f-­t he-­s ustainable-­p otential-­f or-­
bioenergy-­production-­1509.pdf. Accessed on July 17, 2023.
Pillay, A. E., Elkadi, M., & Fok, S. C. (2008). Biofuels – Bane or blessing? Research Jounal of
Chemistry and Environment, 12(2), 5–6.
18 S. E. Haque and T. R. Rafi

Przywara, M., Przywara, R., Zapata, W., & Opalinski, I. (2023). Mechanical properties of solid
biomass as affected by moisture content. AgriEngineering, 5(3), 1118–1135. Available at:
https://doi.org/10.3390/agriengineering5030071
Raghavendra, H., Mishra, S., Upashe, S. P., & Floriano, J. F. (2019). Research and production of
second-generation biofuels. In eBooks (pp. 383–400). John Wiley & Sons, Ltd. Available at:
https://doi.org/10.1002/9781119434436.ch18
Rago, Y. P., Mohee, R., & Surroop, D. (2018). A review of thermochemical technologies for the
conversion of waste biomass to biofuel and energy in developing countries. In W. Leal Filho
& D. Surroop (Eds.), The Nexus: Energy, environment and climate change. Green energy and
technology. Springer. https://doi.org/10.1007/978-­3-­319-­63612-­2_8
REN21. (2011). Renewables 2011: Global status report (pp. 13–14). Available at: https://www.
ren21.net/wp-­content/uploads/2019/05/GSR2011_Full-­Report_English.pdf. Accessed on July
17, 2023.
Rodionova, M. V., Puodyal, R. S., Tiwari, I., Voloshin, R. A., Zharmukhamedov, S. K., Nam,
H. G., et al. (2017). Biofuel production: Challenges and opportunities. International
Journal of Hydrogen Energy, 42(12), 8450–8461. Available at: https://doi.org/10.1016/j.
ijhydene.2016.11.125
Rosillo-Calle, F. (2022). New insights into biomass and biofuels in rapidly changing energy sce-
nario. Energies, 15(18). Available at: https://doi.org/10.3390/en15186664
S&TCI. (2006). Second generation biofuels. A review from a market barrier perspective. https://
task39.ieabioenergy.com/wp-­content/uploads/sites/37/2013/05/Second-­generation-­biofuels-­
A-­review-­from-­a-­market-­barrier-­perspective.pdf
Sajid, Z., Da Silva, M. A. B., & Danial, S. N. (2021). Historical analysis of the role of governance
systems in the sustainable development of biofuels in Brazil and the United States of America
(USA). Sustainability, 13(12). https://doi.org/10.3390/su13126881
Saravankumar, P. T., Suresh, V., Vijayan, V., et al. (2019). Ecological effect of corn oil biofuel with
SiO2 nano-additives. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects,
41(23), 2845–2852. https://doi.org/10.1080/15567036.2019.1576079
Sekoai, P. T., & Yoro, K. O. (2016). Biofuel development initiatives in Sub-Saharan Africa:
Opportunities and challenges. Climate, 4(2). https://doi.org/10.3390/cli4020033
Shokravi, H., Heidarrezaei, M., Shokravi, Z., Ong, H.C., Lau, W.J., et al. (2022). Fourth genera-
tion biofuel from genetically modified algal biomass for bioeconomic development, Journal of
Biotechnology, 360, 23–36, https://doi.org/10.1016/j.jbiotec.2022.10.010
Silvestri, L. C. (2008). The bitter sweet promise of biofuels sweet for few, bitter for many: A study
case of Honduras Luciana, environmental science, agricultural and food sciences. https://www.
researchgate.net/publication/27825584_The_Bitter_Sweet_Promise_of_Biofuels_-_Sweet_
for_few_bitter_for_many_A_sudy_case_of_Honduras. Accessed January 23, 2024.
Simya, O. K., Nair, P. R., & Ashok, A. M. (2018). Engineered nanomaterials for energy applica-
tions. In Elsevier eBooks (pp. 751–767). https://doi.org/10.1016/b978-­0-­12-­813351-­4.00043-­2
Su, Y., Zhang, P., & Su, Y.-Q. (2015). An overview of biofuels policies and industrialization in the
major biofuel producing countries. Renewable & Sustainable Energy Reviews, 50, 991–1003.
https://doi.org/10.1016/j.rser.2015.04.032
Suali, E., & Suali, L. (2023). Impact assessment of global biofuel regulations and policies on
biodiversity. In Elsevier eBooks (pp. 137–161). https://doi.org/10.1016/b978-­0-­323-­91159-
­7.00012-­6
Subramaniam, Y., & Masron, T. A. (2021). The Impact of economic globalization on biofuel
in developing countries. Energy Conversion and Management, X, 10, 100064. https://doi.
org/10.1016/j.ecmx.2020.100064.
Surriya, O., Saleem, S. S., Waqar, K., Kazi, A. G., & Öztürk, M. (2015). Bio-fuels:
A blessing in disguise. In Springer eBooks (pp. 11–54). Available at: https://doi.
org/10.1007/978-­94-­007-­7887-­0_2
Is Biomass Energy a Boon or Bane for Society: A Comprehensive Analysis 19

Szwaja, S., et al. (2013). Sewage sludge producer gas enriched with methane as a fuel to a
spark ignited engine. Fuel Processing Technology, 110, 160–166. https://doi.org/10.1016/j.
fuproc.2012.12.008
TET. (2020). India aims to reduce carbon footprint by 30–35%: PM Narendra Modi. https://
economictimes.indiatimes.com/industry/energy/power/india-­has-­set-­target-­of-­cutting-­carbon-­
footprint-­by-­30-­35-­narendra-­modi/articleshow/79336625.cms?from=mdr. Accessed on July
17, 2023.
TET (The Economic Times). (2018). India’s oil import bill to jump by 25% in FY18. https://
economictimes.indiatimes.com/industry/energy/oil-­gas/indias-­oil-­import-­bill-­to-­jump-­by-­25-­
in-­fy18/articleshow/63464408.cms?utm_source=contentofinterest&utm_medium=text&utm_
campaign=cppst. Accessed on July 17, 2023.
Tripathi, N., Hills, C. D., Singh, R. S., & Atkinson, C. J. (2019). Biomass waste utilisation in low-­
carbon products: Harnessing a major potential resource. npj Climate and Atmospheric Science,
2(1). https://doi.org/10.1038/s41612-­019-­0093-­5
Uddin, M. N., Taweekun, J., Techato, K., Rahman, M. A., Mafijur, M., et al. (2019). Sustainable
biomass as an alternative energy source: Bangladesh perspective. Energy Procedia, 160,
648–654. https://doi.org/10.1016/j.egypro.2019.02.217
Ullah, K., Sharma, V. K., Dhingra, S., Braccio, G., Ahmad, M., et al. (2015). Assessing the ligno-
cellulosic biomass resources potential in developing countries: A critical review. Renewable
& Sustainable Energy Reviews, 51, 682–698. Available at: https://doi.org/10.1016/j.
rser.2015.06.044
UNCTAD. (2016). Advanced biofuels set to play key role in developing countries. https://unctad.
org/news/advanced-­biofuels-­set-­play-­key-­role-­developing-­countries
UNCTAD. (2023). Trade Development Report 2023. Growth, debt, and climate: realigning the
global financial architecture. https://unctad.org/system/files/official-document/tdr2023_en.pdf.
Accessed January 24, 2024.
U.S. Energy Information Administration. (2022). Biofuels explained. https://www.eia.gov/energy-
explained/biofuels/#:~:text=In%202021%2C%20about%2017.5%20billion,and%20net%20
exports%20of%20biofuels. Accessed January 24, 2024.
USDA. (2007). Rising biofuels prices rising biofuels prices – Blessing or curse for food secu-
rity? USDA Global Conference on Agricultural Biofuels. https://www.ars.usda.gov/meetings/
Biofuel2007/presentations/Econ%20Outlook/Gurkan.pdf. Accessed on July 10, 2023.
USDA. (2019). Biofuels annual China will miss E10 by 2020 goal by wide margin. GAIN Report
Number: CH19047. https://apps.fas.usda.gov/newgainapi/api/report/downloadreportbyfile
name?filename=Biofuels%20Annual_Beijing_China%20-­%20Peoples%20Republic%20
of_8-­9-­2019.pdf. Accessed on July 17, 2023.
USDA. (2022). Biofuel mandates in the EU by Member State – 2022. Report Number: E42022-0044.
https://apps.fas.usda.gov/newgainapi/api/Report/DownloadReportByFileName?fileName=Bio
fuel%20Mandates%20in%20the%20EU%20by%20Member%20State%20-­%202022_Berlin_
European%20Union_E42022-­0044.pdf. Accessed on July 17, 2023.
USDE (United States Department of Energy). (n.d.). Biofuels & greenhouse gas emissions: Myths
versus facts. https://www.energy.gov/articles/biofuels-­greenhouse-­gas-­emissions-­myths-­
versus-­facts-­0. Accessed on July 16, 2023.
USEPA. (2023a). Global greenhouse gas emissions data. Available at: https://www.epa.gov/
ghgemissions/global-­greenhouse-­gas-­emissions-­data. Accessed: July 9, 2023.
USEPA. (2023b). Economics of biofuels. Available at: https://www.epa.gov/environmental-­
economics/economics-­biofuels. Accessed: July 10, 2023.
Usmani, R. A. (2020). Potential for energy and biofuel from biomass in India. Renewable Energy,
155, 921–930. Available at: https://doi.org/10.1016/j.renene.2020.03.146. Accessed: July
10, 2023.
Van Ruijven, B., De Cian, E., & Wing, I. S. (2019). Amplification of future energy demand
growth due to climate change. Nature Communications, 10(1). https://doi.org/10.1038/
s41467-­019-­10399-­3
20 S. E. Haque and T. R. Rafi

Weng, Y., Chang, S., Cai, W., & Wang, C. (2019). Exploring the impacts of biofuel expansion on
land use change and food security based on a land explicit CGE model: A case study of China.
Applied Energy, 236, 514–525. https://doi.org/10.1016/j.apenergy.2018.12.024
Zhang, Q., Watanabe, M., & Lin, T. (2020). Rural biomass energy 2020: People’s Republic of
China. Available at: https://www.adb.org/sites/default/files/publication/27997/rural-­biomass-­
energy-­2020.pdf. Accessed: June 8, 2023.
Zhou, C., & Wang, Y. (2020). Recent progress in the conversion of biomass wastes into functional
materials for value-added applications. Science and Technology of Advanced Materials, 21(1),
787–804. https://doi.org/10.1080/14686996.2020.1848213
Ziolkowska, J. R. (2020). Chapter 1 – Biofuels technologies: An overview of feedstocks, pro-
cesses, and technologies. In J. Ren, A. Scipioni, A. Manzardo, & H. Liang (Eds.), Biofuels
for a more sustainable future (pp. 1–19). Elsevier. https://doi.org/10.1016/B978-­0-­12-­815581-­
3.00001-­4
Approach to Reduce Agricultural Waste
via Sustainable Agricultural Practices

Prasann Kumar, Amit Raj, and Vantipalli Aravind Kumar

1 Introduction

Agricultural waste, encompassing residues from farming activities and their impli-
cations for ecosystems, is a pivotal area of scientific interest (Sharma & Bhardwaj,
2017). This composite category engulfs diverse constituents, spanning agricultural
by-products such as stems, leaves, husks, livestock dung, and urine. It further
encompasses non-biodegradable components such as plastic mulch, herbicides, and
fertilisers, introducing complexities to the agricultural waste domain. The variation
in agricultural practices, crop types, and waste management methodologies engen-
ders heterogeneous patterns in agricultural waste generation and composition.
Notably, regions with pronounced livestock production grapple with the consequen-
tial waste produced by such activities. Conversely, the residues of cereal crops such
as wheat and maize constitute a notable share of this waste stream (Koul et al.,
2022). Failure to effectively manage agricultural waste may have deleterious eco-
logical ramifications, such as atmospheric pollution resulting from the incineration
of crop remnants or water contamination stemming from the discharge of animal
waste into aquatic systems. Likewise, the inadvertent disposal of non-biodegradable
materials, including plastic mulch and pesticide containers, poses additional envi-
ronmental threats (Kumawat et al., 2022). Minimising agricultural waste assumes
paramount significance for a multitude of reasons. First, it mitigates the ecological
toll exacted by farming activities. Second, it fosters the conservation of finite
resources such as land and water. Moreover, reducing agricultural waste holds
potential economic advantages, including cost savings in waste disposal and creat-
ing new revenue streams through repurposing by-products for agricultural use. As
part of the broader objectives of sustainable agriculture, curtailing food waste is a

P. Kumar (*) · A. Raj · V. A. Kumar


Department of Agronomy, School of Agriculture, Lovely Professional University,
Phagwara, Punjab, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 21


A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_2
22 P. Kumar et al.

critical aspiration facilitated by conservation tillage, integrated pest management


(IPM), and composting (Wato et al., 2020). Sustainable farming practices are a
linchpin for maintaining soil health, safeguarding natural resources, and nurturing
biological diversity and resilience (Lal, 2015). These practices, which have gar-
nered increased attention, hold promise in enhancing the overall quality of life by
simultaneously addressing various domains, including biodiversity, human health,
and community cohesion (Umesha et al., 2018). The ethos of sustainable agriculture
is founded on several guiding principles:
Resource Conservation: The heart of sustainable agriculture lies in conserving and
safeguarding natural resources such as soil, biodiversity, and wildlife. This
approach aims to curtail reliance on finite resources such as water and fossil fuels
(Newman & Jennings, 2012).
Minimisation of Non-biodegradable Utilisation: Encouraging the judicious use of
non-biodegradable materials and promoting their reuse and recycling are central
tenets of sustainable agricultural philosophy.
Biodiversity Enhancement: The core of sustainable agriculture entails bolstering
biodiversity and ecological resilience by cultivating a diverse array of crop and
livestock species (Giller et al., 1997).
Natural Pest Control: Sustainable agricultural practices seek to reduce pesticide
dependence while fostering natural pest control through crop rotation, biological
management, and cultural practices (Singh, 2021).
Economic Empowerment: A pivotal facet of sustainable agriculture is strengthening
local economies through enhanced market access for farmers and skill develop-
ment via training and education (Kumar et al., 2015).
In the context of sustainable agriculture, food waste reduction stands as a pri-
mary objective. Practices such as composting enrich soil nutrients by breaking
down organic matter and contributing to soil health and fertility (Ayilara et al.,
2020). Conservation tillage, another sustainable practice, minimises soil distur-
bance and erosion, reducing the need for fossil fuels and mechanical labour while
preserving soil moisture and nutrient levels. Integrated pest management (IPM) is a
complementary approach that reduces pesticide reliance and encourages natural
pest control methods (Deguine et al., 2021). The endeavours to mitigate food waste
bear socio-environmental-economic fruits, reducing waste, preserving the environ-
ment, and strengthening local economies through sustainable farming practices. By
adopting these practices, farmers contribute to sustaining soil health, conserving
natural resources, and fostering ecological diversity and resilience.

2 Reduce Overproduction and Food Waste

With the escalating global population and the subsequent rise in food consumption,
urgent measures are imperative to address the issues of food waste and overproduc-
tion. Squandered food holds significant status as a contributor to greenhouse gas
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 23

emissions and a squandering of valuable resources, entailing substantial economic


and environmental repercussions. The mitigation of overproduction and food waste
can be achieved through multifaceted approaches, several of which are expounded
upon in this chapter (Rohini et al., 2020). These encompass refined crop planning
and demand projection, relaxed aesthetic standards for produce, amplified access to
nutritious sustenance for marginalised populations, initiatives targeting food waste
reduction, and comprehensive consumer education (Table 1).

Table 1 Agricultural waste and its sources of production and disadvantages


Agricultural waste Sources of production Disadvantages
Crop residue Harvested plant parts Soil erosion, nutrient loss
Fruit peel Fruit processing and consumption Landfill contribution, methane
emission
Vegetable trimmings Vegetable processing and Waste generation, landfill impact
consumption
Manure Animal farming and livestock Water pollution, odour, greenhouse
gases
Husks and shells Grain processing Slow decomposition, disposal
challenge
Sawdust Wood processing and carpentry Fire hazard, poor soil structure
Eggshells Poultry farming Slow decomposition, limited uses
Coffee grounds Coffee processing and consumption Landfill impact, waste of organic
matter
Tea leaves Tea processing and consumption Slow decomposition, limited uses
Cotton residue Cotton farming and processing Pesticide residues, soil degradation
Aquaculture waste Fish farming and processing Water pollution, disease spread
Stalks and stems Plant fibre processing Slow decomposition, limited uses
Nutshells Nut processing and consumption Slow decomposition, limited uses
Grass clippings Lawn and garden maintenance Rapid decomposition, waste in
landfills
Food scraps Household and commercial Methane emission, resource waste
kitchens
Leather scraps Leather industry by-products Poor biodegradability, pollution
potential
Feather waste Poultry processing and feather Limited uses, waste disposal
industries challenge
Fish scales Fish processing and consumption Limited uses, odour
Shellfish shells Shellfish processing and Slow decomposition, disposal
consumption challenge
Brewery waste Beer production Odour, waste generation
Winery waste Wine production Odour, waste generation
Peat moss residue Peat extraction and horticulture Habitat destruction, carbon
emissions
Olive pits Olive oil production Slow decomposition, limited uses
Cheese whey Cheese production Water pollution, unpleasant odour
(continued)
24 P. Kumar et al.

Table 1 (continued)
Agricultural waste Sources of production Disadvantages
Silk waste Silk production Poor biodegradability, limited uses
Nutrient-rich sludge Wastewater treatment Water pollution, disposal challenge
Crab and lobster Seafood processing Limited uses, disposal challenge
shells
Potato peels Potato processing and consumption Landfill impact, waste of organic
matter
Sawmill residue Wood processing and lumber Waste generation, disposal
industry challenge
Paper pulp residue Paper production Water pollution, habitat destruction
Source: based on the review of literature

2.1 Improved Crop Planning and Demand Forecasting

Enhancing crop planning and demand projection is pivotal in combating food waste
and surplus. Collaborative efforts by farmers and other stakeholders to align produc-
tion with consumer preferences curtail the risk of overproduction and waste
(Kusumowardani et al., 2022). This can be achieved through diverse means, such as
market research, data analysis, and collaboration within the food sector. Data-driven
insights facilitate streamlined production and avert excess. By fostering improved
alignment between production and demand, waste reduction can be achieved, guided
by comprehensive market insight and consumer interaction (Flanagan et al., 2019).

2.2 Relaxed Cosmetic Standards for Produce

The reduction of food waste can also be actualised by redefining the criteria governing
the aesthetic attributes of produce. Perfectly edible fruits and vegetables often meet
disposal due to cosmetic non-conformities with consumer expectations. The relax-
ation of such standards can constrict waste while fostering the consumption of other-
wise discarded produce (Gunders & Bloom, 2017). Current stringent guidelines,
particularly affecting fresh produce, induce unwarranted disposals. Easing these stan-
dards not only curtails waste but also bolsters supply, as demonstrated by the efficacy
of this strategy in certain supermarket implementations (Young et al., 2018).

2.3 Enhanced Food Distribution and Accessibility


for Underserved Communities

Distribution centres emerge as prominent sources of food waste within the food sec-
tor. Strengthening distribution networks and widening access to sustenance for
resource-limited communities mitigate waste while enhancing food availability (Li
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 25

et al., 2020). Ineffectual distribution systems and constrained accessibility contrib-


ute notably to global food waste. Amplifying linkages connecting producers and
consumers through measures such as infrastructure development and promotion of
local grocery outlets reduces waste. The resulting convenience in acquiring and
consuming fresh, nourishing food fosters community health while minimising food
waste (Alattar et al., 2020).

2.4 Food Waste Reduction Campaigns


and Consumer Education

Raising awareness and educating consumers about the implications of food waste
can catalyse change at individual and collective levels. Empowering individuals
with knowledge about practices such as composting, portion control, and environ-
mental stewardship curtails food wastage. Equipping consumers with information
and actionable insights motivates adopting more sustainable behaviours. Social
media campaigns, educational programmes, and community engagement platforms
offer viable avenues for achieving this transformation. By furnishing individuals
with informed decision-making tools, food waste reduction and promoting sustain-
able food systems can be advanced (Rohini et al., 2020).

3 Sustainable Soil Management

Sustainable soil management is centred on maintaining and enhancing soil health


while concurrently preserving the enduring productivity of agricultural land. Soil is
a vital cornerstone for agriculture, housing essential minerals, and organic matter
indispensable for optimal plant growth. Implementing sustainable soil management
practices offers manifold benefits to farmers and land stewards, encompassing
heightened soil fertility, erosion reduction, water conservation, and biodiversity
enhancement. In this discourse, we delve into several critical practices of sustain-
able soil management (Lal et al., 2011):

3.1 Crop Rotation and Diversification

The age-old practice of crop rotation involves periodic alteration of crops planted in
a given field. This practice yields multifaceted advantages for soil health. By diver-
sifying crops, soil fertility is conserved, given the distinct nutrient requirements and
impacts of various crops on the soil. Additionally, crop rotation mitigates disease
and pest proliferation by depriving them of their preferred food sources during
26 P. Kumar et al.

non-­cultivation periods. Crop diversification, involving the simultaneous cultivation


of multiple crops within a single field, augments soil health, diminishes susceptibil-
ity to pests and diseases, and fosters resilience (Fanadzo et al., 2018).

3.2 Cover Cropping and Green Manures

Incorporating non-cash crops, such as legumes or grasses, as cover or green manures


emerges as an effective strategy for sustaining soil health and bolstering crop yields.
Cover crops shield the soil, elevate organic matter content, and even deter weed
growth (Benincasa et al., 2017). They enhance soil fertility through atmospheric
nitrogen fixation, subsequently released in a usable form. Green manures, desig-
nated plants grown for incorporation into the soil as organic matter sources, foster
soil structure and fertility. These practices diminish reliance on synthetic fertilisers,
augment water retention capacity, and amplify biodiversity (Meena et al., 2018).

3.3 No-Till or Reduced Tillage Farming

Soil manipulation through mechanical tillage, while aiding in weed control and land
preparation, harbours the potential for soil loss, nutrient depletion, and organic mat-
ter degradation (Bhattacharyya et al., 2015). No-till or reduced tillage farming,
eschewing, or limiting soil disturbance during cultivation constitutes an alternative
approach. This technique reduces soil erosion, enhances soil structure, and elevates
soil organic matter content. Water conservation and erosion prevention are added
advantages, alongside the potential for carbon sequestration. Retaining soil micro-
organisms bolsters soil biodiversity while lowering fuel and labour expenses that
accrue from avoiding traditional tillage (Somasundaram et al., 2020).

3.4 Integrated Pest Management (IPM)

Integrated pest management entails a multifaceted approach, encompassing various


techniques to mitigate pest impact, thereby minimising reliance on chemical pesti-
cides. IPM integrates practices such as crop rotation, biological control, cultural
strategies, and selective chemical application. This approach curbs insect damage
while minimising unintended environmental consequences and collateral damage to
non-target organisms. IPM contributes to human and animal safety by diminishing
pesticide exposure, retarding the emergence of resistant pests, and preserving eco-
logical balance (Nawaz et al., 2019).
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 27

3.5 Precision Agriculture and Targeted Fertiliser Application

Precision agriculture harnesses contemporary tools to optimise resource utilisation,


including water and fertilisers. This entails real-time monitoring of soil moisture
and nutrient levels, precise mapping of fields for input application, and computer
modelling to optimise resource allocation (Roy & George, 2020). Precision agricul-
ture entails varied strategies such as variable-rate fertiliser delivery, soil mapping,
and remote sensing to monitor crop health and nutrient demands (Khanal et al.,
2017). This approach conserves water quality, curtails greenhouse gas emissions,
amplifies crop yields, and reduces costs by maximising input efficiency (Shah &
Wu, 2019). Careful fertiliser application mitigates nutrient runoff and leaching, pro-
tecting ecosystems. By targeting fertiliser application only where required, farmers
simultaneously enhance profitability and environmental stewardship (Table 2).

Table 2 List of sustainable soil management practices and their description


Sustainable soil
management practices Description Benefits
Cover cropping Planting cover crops between Erosion prevention, soil fertility
main crops improvement
Crop rotation Alternating crops in a field over Pest and disease control, nutrient
time management
No-till farming Planting crops without Reduced erosion, improved soil
ploughing or tilling structure
Mulching Applying organic or synthetic Moisture retention, weed
mulch to soil suppression
Composting Decomposing organic matter Improved soil structure, nutrient
into nutrient-rich soil availability
Organic matter addition Adding compost, manure, or Soil fertility enhancement,
plant residues water-holding capacity
Reduced chemical inputs Minimising synthetic fertilisers Improved soil and water quality
and pesticides
Agroforestry Integrating trees and crops in Improved soil structure, carbon
the same system sequestration
Integrated nutrient Combining organic and Balanced soil fertility, reduced
management inorganic nutrient sources nutrient runoff
Green manuring Incorporating green plant Nitrogen fixation, nutrient
material into soil enrichment
Conservation tillage Reducing the intensity of tillage Soil erosion prevention, moisture
operations conservation
Windbreak establishment Planting trees or shrubs to Erosion prevention, microclimate
shield from wind improvement
Nutrient cycling Recycling nutrients within Reduced nutrient loss, improved
agroecosystems sustainability
Biochar application Adding charcoal-like substance Carbon sequestration, improved
to soil soil fertility
(continued)
28 P. Kumar et al.

Table 2 (continued)
Sustainable soil
management practices Description Benefits
Subsoil manuring Adding organic materials to Improved root growth, enhanced
subsoil nutrient uptake
Terracing Creating level platforms on Erosion prevention increased
steep slopes arable land
Precision agriculture Using technology to apply Reduced waste, improved nutrient
inputs accurately use efficiency
Crop residue management Retaining crop residues on the Improved organic matter,
soil surface moisture retention
Alley cropping Planting crops between alleys Soil protection, enhanced
of trees agroecosystem diversity
Vermicomposting Composting using earthworms Nutrient enrichment, improved
soil structure
Soil aeration Loosening soil to improve air Enhanced root growth, soil
circulation structure improvement
Following Leaving fields unplanted for a Weed and pest control, soil
period regeneration
Agroecological practices Designing farming systems Biodiversity promotion, reduced
based on local ecology environmental impact
Soil pH management Adjusting soil pH to optimal Improved nutrient availability,
levels enhanced plant growth
Nutrient management Developing strategies for Efficient fertiliser use, reduced
planning nutrient use nutrient loss
Soil erosion control Implementing measures to Soil conservation, maintenance of
prevent erosion topsoil
Soil health assessment Regularly evaluating soil health Informed decision-making, early
problem detection
Carbon farming Focusing on carbon Climate change mitigation,
sequestration in agriculture improved soil health
Agroforestry systems Intercropping trees with crops Biodiversity enhancement,
improved soil structure
Riparian buffer Planting vegetation along water Water quality protection, erosion
establishment bodies prevention
Microbial inoculants Adding beneficial Improved nutrient cycling, disease
microorganisms to the soil suppression
Soil structure Enhancing soil aggregation and Water infiltration improvement,
improvement porosity root penetration
Manure management Proper handling and application Nutrient recycling, reduced water
of manure pollution
Irrigation management Efficient use of water for Reduced water waste, enhanced
irrigation water use efficiency
Agroforestry hedgerows Planting rows of trees and Windbreak, habitat for beneficial
shrubs on boundaries organisms
Strip cropping Alternating different crops in Erosion prevention, pest control
narrow strips
(continued)
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 29

Table 2 (continued)
Sustainable soil
management practices Description Benefits
Soil conservation tillage Reducing soil disturbance while Erosion prevention, moisture
planting conservation
Soil testing and Regularly analysing soil Informed nutrient management,
monitoring properties problem detection
Biofertiliser application Using beneficial Enhanced nutrient availability,
microorganisms as fertilisers reduced reliance on chemicals
Soil amendments Adding materials to improve pH adjustment, nutrient
soil properties enrichment
Erosion-resistant crops Planting crops that provide Erosion prevention, soil
ground cover protection
Sustainable drainage Managing water drainage to Soil conservation reduced
systems prevent erosion waterlogging
Agroecosystem diversity Growing a variety of crops and Pest control, enhanced ecosystem
species services
Natural resource Responsible use of resources Reduced environmental impact,
management such as water and energy sustainable production
Biopesticide application Using natural substances for Reduced chemical use, minimised
pest control environmental impact
Soil conservation Techniques to prevent soil Reduced erosion, improved soil
practices degradation quality
Nutrient fixation Using plants to capture and Enhanced nutrient availability,
fixate nitrogen reduced leaching
Phytoremediation Using plants to remove Soil pollution reduction,
contaminants from soil ecosystem restoration
Green manure cover crops Planting specific crops to Nitrogen fixation, organic matter
improve soil enrichment
Aggregation enhancement Improving soil structure Enhanced water retention, root
through management penetration
Reduced pesticide use Minimising the application of Reduced environmental impact,
chemical pesticides ecosystem health
Silvopasture Combining trees and forage for Improved land-use efficiency,
livestock biodiversity
Intercropping Planting different crops together Space utilisation, reduced pests,
and diseases
Water management Efficiently managing water Reduced water waste, improved
resources plant growth
Indigenous knowledge Incorporating local knowledge Sustainable practices, cultural
integration into practices preservation
Agroforestry alley Intercropping with rows of trees Soil protection, diversified yields
cropping
Soil amendments Adding substances to enhance pH adjustment, nutrient
application soil characteristics enrichment
30 P. Kumar et al.

4 Water Conservation and Management

Sustaining agriculture and safeguarding natural resources necessitate pragmatic


water conservation and management protocols. Escalating populations, shifting cli-
mates, and other factors have precipitated water scarcity in numerous global regions.
Immediate implementation of sustainable water management practices is para-
mount, including drip irrigation, precision irrigation, on-farm water recycling, rain-
water harvesting, deployment of drought-resistant crop varieties, and watershed
management (Russo et al., 2014) (Table 3).

4.1 Drip Irrigation and Precision Irrigation

Drip irrigation and precision irrigation present avenues to curtail water wastage in
agriculture. Drip irrigation, utilising pipelines and emitters, directly delivers water
to plant roots (Manda et al., 2021). This technique can potentially halve water con-
sumption compared to traditional flood irrigation methods. Precision irrigation
employs technology to deliver water judiciously, considering parameters such as
soil moisture, climate, and crop requisites (Adeyemi et al., 2017). With real-time
adjustments in water application rates, precision irrigation can surpass even drip
irrigation’s water savings, promoting long-term water conservation and elevated
crop yields.

4.2 On-Farm Water Recycling and Rainwater Harvesting

Water recycling and rainwater collection offer strategies to minimise freshwater


usage. On-farm water recycling involves reutilising household wastewater for irri-
gation (Kampragou et al., 2011). This approach not only reduces water costs but
also aids in water conservation. Rainwater harvested from rooftops or surfaces dur-
ing rainfall can be stored for subsequent use, which is particularly valuable in
regions with limited precipitation (Sivanappan, 2006).

4.3 Drought-Resistant Crop Varieties and Crop Selection

Adopting drought-resistant crop varieties and diversifying crop types constitute


water shortage management methods. Selective breeding has yielded low-water,
high-yield, drought-resistant crop varieties. These crops enable water savings with-
out compromising harvest yields (Maleksaeidi & Karami, 2013). Thoughtful crop
selection is equally vital; for instance, water-intensive crops such as rice and cotton
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 31

Table 3 Water conservation and management practices and its description and benefits
Water conservation and
management practices Description Benefits
Rainwater harvesting Collecting and storing rainwater Reduced reliance on external
for later use water sources
Drip irrigation Slowly apply water directly to Reduced water waste, improved
plant roots plant health
Xeriscaping Landscaping with drought-­ Reduced water consumption,
resistant plants lower maintenance
Greywater recycling Treating and reusing household Reduced strain on freshwater
wastewater resources
Efficient irrigation Timing irrigation to minimise Improved water use efficiency,
scheduling water loss healthier plants
Mulching Covering soil with materials to Reduced evaporation, weed
retain moisture suppression
Water-efficient appliances Using appliances that require Lower water consumption,
less water reduced bills
Native plant landscaping Using plants adapted to the Reduced water needs,
local climate ecosystem support
Permeable pavements Allowing water to seep through Groundwater recharge, reduced
the pavement runoff
Efficient industrial water use Implementing technologies for Lower operational costs,
water savings environmental benefits
Water-efficient toilets and Using fixtures that reduce water Lower water usage, reduced
faucets consumption water bills
Soil moisture sensors Monitoring soil moisture for Preventing overwatering,
targeted irrigation healthier plants
Artificial wetlands Creating human-made wetlands Water purification, habitat
for water treatment creation
Desalination technologies Removing salt and impurities Increased freshwater
from seawater availability, drought resilience
Water recycling in industry Reusing water for industrial Lower water intake, reduced
processes pollution
Leak detection systems Installing devices to detect and Water conservation, reduced
stop leaks water bills
Efficient watering Using methods that minimise Better plant absorption, reduced
techniques water runoff wastage
Aquifer recharge Injecting treated water into Groundwater replenishment,
underground aquifers future water supply
Water-efficient farming Employing methods that Sustainable agriculture, higher
practices optimise water usage yields
Urban planning for water Designing cities to conserve Reduced flooding, improved
management and manage water water quality
Fog harvesting Collecting water droplets from Alternative water sources in
fog arid regions
Education and awareness Informing people about water Behavioural change,
campaigns conservation community involvement
Water rights and allocation Regulating water distribution Fair allocation, reduced
policies and usage conflicts
(continued)
32 P. Kumar et al.

Table 3 (continued)
Water conservation and
management practices Description Benefits
Rain gardens Landscaping features designed Water conservation, enhanced
to capture rain aesthetics
Smart irrigation controllers Using technology to optimise Reduced water wastage,
irrigation healthier landscapes
Water footprint tracking Monitoring personal and Informed consumption choices,
industrial water usage reduced waste
Soil amendment Adding materials to improve Enhanced soil structure,
management water retention reduced runoff
Harvested fog utilisation Using harvested fog for various Additional water sources in
purposes water-scarce areas
Water pricing strategies Setting prices to reflect water Encouraging efficient use,
scarcity funding conservation
Riverbank filtration Treating river water through Improved water quality, reduced
natural filtration treatment costs
Watershed management Protecting and restoring Enhanced water quality,
watersheds reduced runoff
Agricultural runoff Preventing runoff of chemicals Reduced water pollution,
management from farms healthier ecosystems
Efficient cooling systems Using water-efficient cooling Reduced water consumption,
technologies lower energy use
Cloud seeding Introducing substances to Enhanced precipitation,
encourage rainfall increased water supply
Water-efficient industrial Optimising manufacturing Lower water consumption,
processes processes for water use reduced costs
Constructed wetlands Creating wetland areas for Natural filtration, habitat
water treatment creation
Demand management Reducing water use during peak Avoided shortages, efficient
demand times water usage
Managed aquifer recharge Artificially recharging Groundwater restoration,
groundwater supplies drought resilience
Water recycling in Treating and reusing water in Reduced irrigation needs,
agriculture farming sustainable farming
Water sense certification Labelling water-efficient Consumer awareness, reduced
products water usage
Infiltration basins Capturing and infiltrating Reduced runoff, groundwater
stormwater recharge
Green roofs Planting vegetation on building Improved stormwater
roofs management, insulation
Water-efficient land Planning urban areas with water Reduced runoff, enhanced
development conservation water quality
Water banking Storing excess water for future Drought resilience, water
use availability
Efficient industrial cooling Minimising water use in Reduced water consumption,
cooling processes cost savings
Rainwater purification Treating rainwater for potable Additional water source,
use reduced demand
(continued)
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 33

Table 3 (continued)
Water conservation and
management practices Description Benefits
Floodplain restoration Reestablishing natural Reduced flood risk, improved
floodplain ecosystems water quality
Water-efficient building Designing structures for Reduced consumption, lower
design minimal water use utility bills
Subsurface drip irrigation Delivering water directly to Enhanced plant health, water
plant roots conservation
Water-efficient landscaping Designing outdoor spaces for Lower irrigation needs,
minimal water use improved aesthetics
Water monitoring systems Installing sensors to track water Early pollution detection,
quality informed decisions
Efficient water distribution Minimising water loss in Reduced leakage, increased
networks distribution systems supply efficiency
Desalination powered by Using renewable sources for Reduced environmental impact,
renewable energy desalination sustainable
Water-efficient appliances Installing products that use less Reduced water consumption,
and fixtures water lower bills
Erosion control Implementing measures to Soil conservation, improved
prevent erosion water quality
Water conservation in Applying water-saving Lower water consumption, cost
commercial buildings practices in businesses savings
Efficient industrial cleaning Minimising water use in Reduced water consumption,
processes cleaning operations cost savings
Water-efficient land-use Zoning to promote water-wise Reduced runoff, enhanced
planning development water supply
Water rights trading Allowing trading of water use Flexible water allocation,
allocations efficient usage
Riparian zone management Protecting and restoring Improved water quality, habitat
riparian areas preservation
Water-efficient irrigation Using technologies to optimise Reduced water waste, healthier
systems irrigation landscapes
Climate-responsive water Adjusting strategies based on Effective water use, climate
management climate conditions resilience
Industrial wastewater Treating and purifying Reduced pollution, protection
treatment industrial wastewater of water bodies
Water-efficient behavioural Encouraging responsible water Reduced wastage, community
strategies usage behaviour involvement
Wetland conservation Preserving and restoring Biodiversity protection, water
wetland ecosystems purification
Efficient turf management Using water-wise practices on Reduced irrigation, healthier
lawns turf
Water conservation in Promoting water-saving Student education, reduced
educational settings practices in schools water usage
Monitoring and reducing Identifying and fixing water Reduced water loss, lower costs
leakage system leaks
34 P. Kumar et al.

might not suit regions with scant water supplies. Opting for drought-resistant crops
tailored to the local conditions, such as sorghum and millet, aligns with effective
water management (Laghari et al., 2012).

4.4 Watershed Management and Inter-farm Cooperation

Effective water resource management can be achieved through watershed manage-


ment and inter-farm collaboration. Watershed management encompasses holistic
oversight of groundwater, surface water, and rainwater resources (Syme et al.,
2015). This approach ensures sustainable water utilisation and equitable accessibil-
ity. Collaborative efforts among farmers within a watershed can extend beyond
water management, exemplified by cooperative irrigation infrastructure projects
that enhance water efficiency and diminish total consumption (Liang et al., 2020).
Cooperative irrigation scheduling and data exchange among farmers further amplify
resource efficiency.
In sum, embracing sustainable water management practices remains a pivotal
cornerstone for the resilience of agriculture and the conservation of natural
resources. Through the judicious application of techniques such as drip and preci-
sion irrigation, on-farm water recycling, rainwater harvesting, adoption of drought-­
resistant crops, and collaborative watershed management, farmers can foster optimal
water utilisation and contribute to sustainable agricultural systems.

5 What Are Some Challenges to Implementing These


Practices in Different Regions?

While water conservation and management techniques such as drip irrigation, preci-
sion irrigation, on-farm water recycling, rainwater harvesting, drought-resistant
crop varieties, and watershed management hold promise in addressing water scar-
city and enhancing agricultural productivity, their broad-scale adoption is hindered
by various challenges (Yadav et al., 2022). Noteworthy among these obstacles are
as follows:

5.1 Lack of Awareness and Education

A substantial impediment to implementing these practices lies in farmers’ and poli-


cymakers’ limited knowledge and comprehension. Many might be unaware of the
potential benefits of adopting such methods or lack the requisite expertise to execute
them effectively.
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 35

5.2 High Initial Costs

Deploying water conservation and management practices, especially involving drip


and precision irrigation technologies, can entail considerable initial expenses. These
costs might be prohibitive for farmers with smaller operations or constrained access
to financial resources (Brahmanand & Singh, 2022).

5.3 Limited Resource Access

The availability of water resources, including groundwater, surface water, and rain-
water, might be insufficient in numerous regions, impeding effective implementa-
tion. Scarce water supplies can hinder the adoption of measures such as rainwater
harvesting or water recycling on farms.

5.4 Climate Variability

The erratic nature of climate, encompassing droughts, floods, and extreme weather
events, poses challenges to the proper execution of water conservation and manage-
ment practices. Prolonged drought periods, for example, might undermine the effi-
cacy of drought-resistant crop varieties.

5.5 Policy and Institutional Barriers

The pervasive uptake of water-saving practices can sometimes be obstructed by


institutional and policy barriers. Various factors, including regulatory limitations on
water recycling and disparities in water rights allocation, can curtail farmers’ access
to water resources (Cairns, 2018).

5.6 Cultural and Social Barriers

Cultural and social dynamics can also shape the diffusion of water-saving and man-
agement methods. Integrating novel practices might encounter resistance when they
clash with established farming practices or prevailing social norms (Liehr
et al., 2016).
36 P. Kumar et al.

In pursuing sustainable water use and bolstered agricultural output, addressing


these multifaceted challenges that inhibit the widespread adoption of effective water
conservation and management practices is imperative.

6 Renewable Energy and Biofuels

The contemporary agricultural sector heavily relies on fossil fuels for machinery,
transportation, and fertilisers, yet this reliance contributes to greenhouse gas emis-
sions and the depletion of finite resources. Agriculture must transition towards
renewable energy sources and biofuels to secure long-term viability (Table 4).

6.1 Anaerobic Digestion for Biogas Production

The anaerobic digestion of agricultural waste can yield biogas, a renewable energy
source. Anaerobic digesters break down organic matter such as manure, food scraps,
and residues to produce methane-rich biogas and nutrient-rich digestate. The biogas
can be converted to energy or heat for agricultural operations, while digestate can
serve as an organic fertiliser, thus closing the loop on waste management
(Ardebili, 2020).

6.2 Conversion of Crop Residues and Waste to Biofuels

Bioconversion transforms organic waste and crop residues into usable biofuels such
as ethanol or biodiesel. Processes such as fermentation convert complex organic
materials into simpler chemicals suitable for biofuel production. Ethanol, for
instance, can be derived from plant materials through fermentation, while biodiesel
can be synthesised from the fatty acids present in biomass (Sindhu et al., 2016;
Mandari & Devarai, 2021).

6.3 Renewable Energy Sources for Agricultural Operations

Leveraging renewable energy sources such as solar, wind, and geothermal power
can render agricultural operations sustainable and cost-effective. Solar panels, wind
turbines, and geothermal energy can power machinery, irrigation systems, and other
farm activities, reducing reliance on fossil fuels and mitigating greenhouse gas
emissions (Rahman et al., 2022; Bhattacharjee & Nayak, 2019).
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 37

Table 4 Renewable energy/biofuel its advantages and disadvantages


Renewable energy/
biofuel Advantages Disadvantages Benefits
Solar photovoltaic Abundant, clean Intermittent, Reduced emissions,
energy energy source weather-dependent decentralised power
Wind energy Low operating costs, Intermittent, visual impact Low carbon footprint,
large-scale power job creation
Hydropower Reliable, low operating Environmental impact, Emission-free, energy
costs habitat disruption storage
Biomass energy Utilises organic waste, Emission of pollutants, Reduces waste,
waste reduction land competition potential
carbon-neutrality
Geothermal energy Continuous Location-specific, Reliable, baseload
availability, minimal resource depletion power generation
emissions
Tidal energy Predictable, high High installation costs, Renewable, minimal
energy density environmental impact emissions
Ocean thermal Constant source, Efficiency challenges, Continuous energy,
energy potential desalination infrastructure costs reduced fossil fuels
Concentrated solar Energy storage, high Water use, high upfront Produces electricity
power (CSP) efficiency costs day and night
Bioenergy (organic Uses waste, waste Land competition, Reduced waste,
materials) reduction emissions balanced carbon cycle
Municipal waste to Waste reduction, Air pollution, ash disposal Reduces landfill
energy energy recovery space, generates
power
Ethanol (from Reduced petroleum Land competition, Renewable supports
various feedstocks) use, lower emissions energy-intensive agriculture
Biodiesel (from Lower emissions, Land competition, Renewable, reduced
various feedstocks) existing infrastructure potential impact on food fossil fuel use
prices
Algal biodiesel High oil yield, Energy-intensive Efficient land use,
potential wastewater cultivation, technical carbon-neutral
treatment challenges
Cellulosic ethanol Non-food feedstocks, Complex production, Less competition with
waste reduction potential environmental food crops
impact
Renewable natural Methane capture, Methane leakage, Reduces methane
gas (RNG) versatile applications infrastructure emissions, sustainable
requirements
Biobutanol Higher energy content, Production complexity, Lower emissions,
existing infrastructure lower energy efficiency potential substitute
Syngas (from Versatile, waste Efficiency loss, technical Converts waste,
biomass gasification) utilisation challenges supports circular
economy
Green diesel Similar properties to Production energy Reduced emissions,
(renewable diesel) diesel, lower emissions intensity, feedstock suitable for engines
availability
Hydrothermal Converts wet biomass, Energy-intensive Converts waste,
liquefaction versatile product range challenges in scaling up renewable fuel
Waste vegetable oil Uses waste oil, reduces Quality variability, limited Waste utilisation,
biodiesel environmental impact feedstock lower carbon footprint
38 P. Kumar et al.

6.4 Reduced Reliance on Fossil Fuels

Shifting towards renewable energy sources, including solar, wind, and geothermal
power, offers multifaceted benefits, encompassing the reduction of greenhouse gas
emissions, decreased dependence on fossil fuels, and enhanced energy affordability
(Sen et al., 2016; Soltani et al., 2021). Combining these energy technologies with
sustainable practices such as water conservation and soil health management can
form a comprehensive approach towards sustainable agriculture.
The agriculture industry can achieve a more sustainable future by embracing
renewable energy sources, transitioning to biofuels, and incorporating efficient
waste management practices. Adopting these practices curbs greenhouse gas emis-
sions and reliance on fossil fuels and enhances resource efficiency, soil health, and
environmental stewardship.

7 Reduced Deforestation and Sustainable Grazing

Greenhouse gas emission mitigation, biodiversity preservation, and rural livelihood


enhancement can be significantly advanced through reduced deforestation and sus-
tainable grazing practices. A range of methods and approaches, including silvopas-
ture, enhanced grassland management, agroforestry and intercropping, and forest
conservation, are employed within these practices (Vera et al., 2022).

7.1 Silvopasture for Integrated Land Use

Silvopastoral systems merge tree cultivation, forage production, and animal graz-
ing. Within this integrated system, trees, grasslands, and grazing animals coexist.
Silvopasture systems offer improved animal welfare, enhanced business output, and
healthier ecosystems. Trees aid in moisture retention, erosion reduction, carbon
storage, and providing shelter and food for livestock. Moreover, they foster biodi-
versity by creating habitats for various organisms.
Implementing silvopasture requires meticulous planning to ensure the compati-
bility of trees and forage with grazing animals. Proper tree selection based on graz-
ing tolerance and fodder production capacity is crucial. Careful management of
grazing intensity is needed to prevent overgrazing and secure adequate fodder pro-
duction. Despite challenges, silvopasture has proven effective for sustainable live-
stock production, contributing to climate change mitigation and biodiversity
preservation (Djanibekov et al., 2015).
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 39

7.2 Enhanced Grassland Management

Grasslands are vital ecosystems that store carbon and support diverse flora and
fauna. Overgrazing and land-use changes contribute to grassland degradation.
Effective grassland management practices, including rotational grazing, reducing
stocking rates, enhancing soil fertility, and managing invasive species, can restore
degraded grasslands, promote biodiversity, and enhance ecosystem services. These
practices aid in mitigating carbon sequestration and climate change (Yang
et al., 2019).

7.3 Agroforestry and Nitrogen-Fixing Trees

Agroforestry systems integrate trees into agricultural and livestock operations,


yielding improved productivity, soil fertility, biodiversity, and reduced greenhouse
gas emissions. Intercropping with nitrogen-fixing trees, like legumes, reduces the
need for synthetic fertilisers while enhancing soil fertility. Trees that fix atmospheric
nitrogen make it available to crops, resulting in higher yields, improved soil health,
and reduced fertiliser use.
Establishing agroforestry demands careful management to ensure compatibility
between trees and crops and selecting suitable species for the local environment.
While agroforestry requires a longer-term investment due to trees’ longer establish-
ment time, it effectively enhances ecosystem services and promotes environmen-
tally responsible farming (Akinnifesi et al., 2008).

7.4 Forest Conservation and Sustainable Land Use

Forests are critical in carbon storage, biodiversity maintenance, and ecosystem ser-
vices. However, deforestation and shifting land use threaten their existence.
Protecting forests through conservation, restoration, sustainable management, and
reducing agricultural conversion is essential. Tactics such as protected areas, forest
restoration, and sustainable forest management complement each other in halting
deforestation and safeguarding biodiversity.
Forest conservation requires land-use planning, zoning regulations, and incen-
tives for sustainable practices. Measures such as incentives for sustainable prac-
tices, proper land-use planning, and zoning regulations can deter future deforestation
and promote sustainable land use (Smith et al., 2016).
In conclusion, reduced deforestation and sustainable grazing practices, encom-
passing methods such as silvopasture, enhanced grassland management, agrofor-
estry, and forest conservation, offer multifaceted benefits for mitigating climate
40 P. Kumar et al.

change, biodiversity conservation, and rural development. Despite the challenges


posed by careful planning and local adaptations, these practices represent powerful
tools for fostering environmental stewardship and long-term sustainability.

8 Technological and Policy Innovations

The agricultural sector around the world is having trouble. The industry has adopted
new practices and technology that can promote sustainable food production in
response to the increasing global population, the effects of climate change, and the
scarcity of natural resources. Recent years have seen a proliferation of technology
and legislative advances to mitigate these difficulties. This chapter will discuss a
few of these new developments and how they might encourage environmentally
responsible farming methods.

8.1 Precision Fermentation, Aquaponics, Vertical Farming,


and Other Innovations

Precision fermentation is an emerging technology that can significantly alter the


food production industry. Using this method, microorganisms such as yeast and
bacteria produce proteins, lipids, and other nutrients usable in the food industry
(Ghosh et al., 2022). In contrast to conventional farming, precision fermentation
reduces the need for land and water. Instead, this can be accomplished with minimal
resources in a laboratory setting. Precision fermentation produces food with less
environmental impact than conventional farming methods. The raising of animals
for food, for instance, is one of the leading causes of global warming, deforestation,
and water pollution. Producing food by precision fermentation eliminates the need
for animal farms, reducing these effects (Awuchi et al., 2020). Aquaponics is a new
method of sustainable farming that uses revolutionary technologies. Aquaponics is
a method of growing plants in water that has recently been combined with aquacul-
ture to create a self-sustaining system for cultivating fish and vegetables (Wirza &
Nazir, 2021). The plants get nutrients from the fish excrement, while the fish get
clean water from the plants. There are many advantages of aquaponics over conven-
tional farming. Because water is constantly recycled throughout the system, it
requires less water than conventional farming (AlShrouf, 2017). Since the fish poop
contains all the nutrients the plants require, there is no need for any additional syn-
thetic fertilisers. Aquaponics also eliminates the need for food to be transported at
great distances because it can be practised in urban settings. Another cutting-edge
method that can aid in sustainable farming is vertical farming. Vertical farming is
cultivating plants by stacking them vertically, typically inside. Vertical farming is a
method of producing a high yield from a small footprint by growing plants
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 41

vertically. There are many advantages of vertical farming over conventional farming
methods. One advantage is that water may be reused multiple times, meaning less
water is used than in conventional farming. The usage of pesticides on plants is also
avoided because of the contained nature of the growing conditions. The fact that
vertical farming can be done in cities means less food has to be transported long
distances (Chatterjee et al., 2020).

8.2 Government Policies Promoting Sustainable


Agricultural Practices

These can be as important as technological advances in fostering sustainable agri-


culture. Reducing greenhouse gas emissions, conserving natural resources, and fos-
tering biodiversity are all examples of sustainable agriculture practices that
governments worldwide have begun enacting (Chopra et al., 2022). The Indian gov-
ernment has introduced several initiatives to encourage farmers to adopt environ-
mentally friendly farming methods. Essential programmes and policies include:

8.2.1 National Mission on Sustainable Agriculture (NMSA)

The National Mission on Sustainable Agriculture (NMSA), established in 2010,


promotes sustainable farming methods in India. Promoting farming methods that
can withstand changing weather conditions and a lack of water is central to the
objective (Gupta et al., 2021).

8.2.2 Pradhan Mantri Fasal Bima Yojana (PMFBY)

In the event of crop loss due to natural disasters or other unforeseeable circum-
stances, farmers can turn to the Pradhan Mantri Fasal Bima Yojana (PMFBY), a
crop insurance scheme. The programme inspires farmers to increase their produc-
tivity with environmentally friendly methods.

8.2.3 Soil Health Card Scheme

A programme to issue soil health cards to agriculturalists, the Soil Health Card
Scheme, was initiated in 2015. Advice on how much and what kind of fertilisers and
other inputs should be used to improve the soil’s nutrient content is included on the
cards. The programme encourages environmentally friendly farming methods such
as using natural fertilisers and organic inputs.
42 P. Kumar et al.

8.2.4 Paramparagat Krishi Vikas Yojana (PKVY)

One programme that encourages organic farming in India is the Paramparagat


Krishi Vikas Yojana (PKVY). Farmers who participate in the programme receive
financial support, training, and technical assistance as they transition to organic
farming (Reddy, 2018).

8.2.5 National Agriculture Market

In India, agricultural goods can be bought and sold online through the National
Agriculture Market (eNAM). Because of the platform’s emphasis on efficiency and
openness, the agricultural marketing system is more likely to support environmen-
tally friendly farming methods.

8.2.6 Rashtriya Krishi Vikas Yojana

The Rashtriya Krishi Vikas Yojana (RKVY) programme helps states pay for agri-
cultural improvement initiatives. This plan encourages farmers to switch to more
eco-friendly methods, including agroforestry and conservation farming (Basim
et al., 2022).

8.2.7 Public-Private Partnerships to Fund Agricultural Innovation


and Transition

Funding agricultural innovation and shifting to more sustainable and resilient farm-
ing practices are two areas where public-private partnerships (PPPs) may signifi-
cantly impact. PPPs can catalyse research and development, speed up the adoption
of novel innovations and practices, and leverage funds to have a more significant
impact since they bring together the resources, experience, and viewpoints of both
the public and private sectors (Hermans et al., 2019). PPPs can help with agricul-
tural change and innovation in numerous ways:
1. Funding Research and Development: In order to promote agricultural produc-
tion, sustainability, and resilience to climate change, PPPs can pool resources
from the public and private sectors in order to fund the development and research
of new technology and practices (Smyth et al., 2021).
2. Developing New Markets and Value Chains: By utilising the knowledge and
connections of private sector partners, PPPs may speed up the creation of new
markets and supply chains for environmentally friendly agricultural goods while
guaranteeing farmers a decent return on their efforts.
3. Promoting Knowledge Sharing and Capacity Building: Public-private partner-
ships (PPPs) can improve communication and collaboration between govern-
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 43

ment agencies, businesses, and other organisations involved in the agricultural


value chain. This can aid in the efficient and long-term implementation of novel
methods (Ferroni & Castle, 2011).
4. Strengthening Policy and Regulatory Frameworks: Besides promoting sustain-
able land use, safeguarding biodiversity, and lowering greenhouse gas emis-
sions, PPPs can engage with governments to develop legislative and regulatory
frameworks that promote sustainable agriculture (Furumo & Lambin, 2020).
Several obstacles public-private partnerships in agriculture must overcome
threaten their efficiency and longevity. Among the difficulties are:
1. Conflicting Interests: Many distinct parties with varying goals, interests, and pri-
orities are typically involved in public-private partnerships. Disagreements and
conflicts can arise over things such as ownership of intellectual property, distri-
bution of profits, and management of the organisation as a whole (Heydari
et al., 2020).
2. Unequal Power Dynamics: Private corporations may have an advantage over
public sector partners and small farmers regarding resources, knowledge, and
bargaining strength. This can lead to decision-making, resource distribution, and
power dynamics inequalities.
3. Limited Accountability: It is possible that public-private partnerships are not
held to the same standards of transparency and accountability as government
programmes or business endeavours. This can make it difficult to monitor every-
thing that is going on and hold anyone accountable (Keers & van Fenema, 2018).
4. Sustainability: While public-private partnerships are frequently formed to tackle
certain issues or complete specific projects, they may not be viable. Reasons for
this include economic fluctuations, government instability, and a lack of finan-
cial resources.
5. Equity and Social Justice: Small farmers, women, and indigenous tribes may be
overlooked when public-private partnerships are formed. This can cause dis-
crimination, unequal access to resources, and other forms of social wrongdoing.

9 Social and Economic Dimensions

Agriculture is crucial for many countries’ economies since it provides food and raw
materials for other sectors. However, agriculture can potentially affect ecosystems,
communities, and economies significantly. Recent years have seen a rise in efforts
to connect farmers with consumers, improve agricultural education, and spread
environmentally friendly farming methods. This chapter will explore such initia-
tives’ social and economic facets, touching on farmer education, community-­
supported agriculture cooperatives, financial incentives for environmentally
responsible farming methods, and fair trade.
44 P. Kumar et al.

Farmer Education, Community-Supported Agriculture, and Cooperatives:


Educating Farmers Can Only Promote Sustainable and Environmentally
Friendly Farming Practices Farmers must consider the short-term results of their
actions and the long-term effects on the ecosystem. In addition, it is essential to
educate farmers on the latest sustainable agriculture practices, tools, and methods.
Extension services, training programmes, and workshops are just a few methods to
educate farmers (Piñeiro et al., 2020). Farmers can get help from the government or
non-profit organisations that offer extension services. Several nations have set up
extension programmes to give farmers access to sustainable agricultural education
and resources. For instance, sustainable agriculture extension programmes receive
money from the US Department of Agriculture and the National Institute of Food
and Agriculture (NIFA) (Osmond et al., 2012). Educating farmers can also be
accomplished with training programmes and workshops. Governmental organisa-
tions, NGOs, and private businesses can all run such initiatives. In Zambia, for
instance, a conservation agricultural project programme instructs farmers in conser-
vation agricultural methods, including limited tillage and crop rotation (Umar et al.,
2011). Sustainable farming practices and farmer prosperity can be advanced through
cooperatives and community-supported agriculture (CSA). In a CSA, customers
and farmers work together, with customers purchasing a crop share in advance.
Shortening the distance between producers and customers benefits farmers by
ensuring a steady market, cutting down on marketing expenses, and encouraging
environmentally friendly farming methods (Pingali et al., 2005). Farmers may also
benefit from cooperatives or other similar models. Cooperatives are businesses in
which the members own and run the business collectively and distribute the earn-
ings and other benefits to the members. Farmers who join a cooperative can get
access to new markets, sources of capital, and knowledge. They also help farmers
speak with one voice regarding government decisions.

Incentives for Farmers to Adopt Sustainable and Eco-Friendly Practices Farmers


might be enticed to adopt more environmentally friendly practices using financial
incentives. Incentive programmes can take numerous forms, including monetary
rewards, expert advice, and public acclaim. The use of financial incentives to
encourage sustainable agriculture practices is widespread. Farmers who implement
environmentally friendly methods may be eligible for government subsidies, grants,
and tax incentives (Soundarrajan & Vivek, 2016). The Conservation Reserve
Programme of the United States Department of Agriculture, for instance, offers
monetary incentives to farmers who transform unproductive farmland into forested
or grassy conservation reserves (Searchinger et al., 2020). Farmers eager to imple-
ment eco-friendly and sustainable practices can receive technical assistance.
Training, expert opinion, and access to data on cutting-edge methods are all exam-
ples of technical support. In the United States, for instance, farmers interested in
sustainable agriculture can get help from the Sustainable Agriculture Research and
Education (SARE) programme. Sustainable farming practices can be encouraged
through recognition programmes as well. Farmers who attempt to adopt sustainable
farming methods are sometimes recognised through recognition programmes.
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 45

Farmers in the United States, for instance, can earn rewards for their work protect-
ing soil, water, and animal habitats under the Conservation Management Programme
(Taylor & Van Grieken, 2015).

10 Conclusions and Future Outlook

Profound challenges in food security, ecological integrity, and societal equity influ-
ence the global food and agricultural framework. The prevalent unsustainable meth-
odologies within this system have led to soil quality degradation, biodiversity
depletion, and escalation of global climate shifts. Many individuals experience
undernourishment and malnourishment due to disparities inherent in the food sup-
ply chain, while others grapple with obesity and ailments tied to dietary patterns.
Despite these formidable impediments, promising prospects come to the fore.
Precision agriculture, genetic modifications, and data-driven analytics have paved
the way for ecologically sound food production. A surge in consciousness among
consumers and governance bodies is driving demand for equitable and ecologically
mindful food systems. This surging awareness has spurred interest in organic,
locally sourced, and fair-trade produce, potentially motivating farmers to embrace
more sustainable methodologies. A recalibration is essential across policies, tech-
nologies, and societal norms to achieve sustainable food and agricultural outcomes.
A pivotal concern lies in harmonising the endeavours of the various stakeholders
within the food system. Farmers, governments, consumers, and researchers all
champion sustainability, yet their undertakings often lack synchronisation. Another
challenge lies in the inadequate funding of sustainable agriculture. Numerous farm-
ers, particularly those in developing nations operating on a smaller scale, encounter
difficulties in accessing financial resources and investments, a predicament that
impedes the adoption of sustainable practices. The hesitancy of investors and finan-
cial institutions to engage with sustainable agriculture stems from perceived risks
and uncertainties. Addressing these multifaceted challenges necessitates a systemic
overhaul towards sustainable food and agriculture. This endeavour mandates a com-
prehensive approach spanning the entire food chain, from cultivation to consump-
tion. The collaborative actions of food system participants should prioritise the
well-being of society, the environment, and the economy. In this context, agroecol-
ogy emerges as a prospective solution. By considering the holistic agroecosystem,
agroecology can bolster food output while enhancing biodiversity, soil vitality, and
food self-determination. It empowers farmers and fosters community-centred food
systems, thereby nurturing social equity. Inspired by natural ecosystems, circular
agriculture operates by cycling nutrients and resources, championing resource effi-
ciency, resilience, and waste curtailment. In summary, the journey towards sustain-
able food and agriculture necessitates comprehensive transformations at a systemic
level, necessitating synchronised efforts, innovative approaches, and a resolute
commitment to balance ecological, societal, and economic imperatives.
46 P. Kumar et al.

References

Adeyemi, O., Grove, I., Peets, S., & Norton, T. (2017). Advanced monitoring and management
systems for improving sustainability in precision irrigation. Sustainability, 9(3), 353. https://
doi.org/10.3390/su9030353
Akinnifesi, F. K., Sileshi, G., Ajayi, O. C., Chirwa, P. W., Kwesiga, F. R., & Harawa, R. (2008).
Contributions of agroforestry research and development to the livelihood of smallholder farm-
ers in Southern Africa: 2. Fruit, medicinal, fuelwood and fodder tree systems. Agricultural
Journal, 3(1), 76–88.
Alattar, M. A., DeLaney, J., Morse, J. L., & Nielsen-Pincus, M. (2020). Food waste knowledge,
attitudes, and behavioural intentions among university students. Journal of Agriculture, Food
Systems, and Community Development, 9(3). https://doi.org/10.5304/jafscd.2020.093.004
AlShrouf, A. (2017). Hydroponics, aeroponic and aquaponic as compared with conventional farm-
ing. American Scientific Research Journal for Engineering, Technology and Sciences, 27(1),
247–255.
Ardebili, S. M. S. (2020). Green electricity generation potential from biogas produced by anaero-
bic digestion of farm animal waste and agriculture residues in Iran. Renewable Energy, 154,
29–37. https://doi.org/10.1016/j.renene.2020.02.102
Awuchi, C. G., Awuchi, C. G., Ukpe, A. E., Asoegwu, C. R., Uyo, C. N., & Ngoka, K. E. (2020).
Environmental impacts of food and agricultural production: A systematic review. European
Academic Research, 8(2), 1120–1135.
Ayilara, M. S., Olanrewaju, O. S., Babalola, O. O., & Odeyemi, O. (2020). Waste management
through composting: Challenges and potentials. Sustainability, 12(11), 4456. https://doi.
org/10.3390/su12114456
Basim, N. M. A., Rajarajan, S., & Ramkumar, P. (2022). Revitalised sustainable agriculture through
ecosystem management and policy interventions: Evidences from India. In Social morphology,
human welfare, and sustainability (pp. 577–602). Springer International Publishing.
Benincasa, P., Tosti, G., Guiducci, M., Farneselli, M., & Tei, F. (2017). Crop rotation as a system
approach for soil fertility management in vegetables. In Advances in research on fertilization
management of vegetable crops (pp. 115–148). Springer.
Bhattacharjee, S., & Nayak, P. K. (2019). PV-pumped energy storage option for convalescing per-
formance of hydroelectric station under declining precipitation trend. Renewable Energy, 135,
288–302. https://doi.org/10.1016/j.renene.2018.12.021
Bhattacharyya, R., Ghosh, B. N., Mishra, P. K., Mandal, B., Rao, C. S., Sarkar, D., et al. (2015).
Soil degradation in India: Challenges and potential solutions. Sustainability, 7(4), 3528–3570.
https://doi.org/10.3390/su7043528
Brahmanand, P. S., & Singh, A. K. (2022). Precision irrigation water management-current status,
scope and challenges. Indian Journal of Fertilisers, 18, 372–380.
Cairns, M. R. (2018). Metering water: Analyzing the concurrent pressures of conservation, sus-
tainability, health impact, and equity in use. World Development, 110, 411–421. https://doi.
org/10.1016/j.worlddev.2018.06.001
Chatterjee, A., Debnath, S., & Pal, H. (2020). Implication of urban agriculture and vertical farming
for future sustainability. In Urban horticulture-Necessity of the future. IntechOpen.
Chopra, R., Magazzino, C., Shah, M. I., Sharma, G. D., Rao, A., & Shahzad, U. (2022). The role
of renewable energy and natural resources for sustainable agriculture in ASEAN countries:
Do carbon emissions and deforestation affect agriculture productivity? Resources Policy, 76,
102578. https://doi.org/10.1016/j.resourpol.2022.102578
Deguine, J. P., Aubertot, J. N., Flor, R. J., Lescourret, F., Wyckhuys, K. A., & Ratnadass,
A. (2021). Integrated pest management: Good intentions, hard realities. A review. Agronomy
for Sustainable Development, 41(3), 38. https://doi.org/10.1007/s13593-­021-­00689-­w
Djanibekov, U., Dzhakypbekova, K., Chamberlain, J., Weyerhaeuser, H., Zomer, R., Villamor, G.,
& Xu, J. (2015). Agroforestry for landscape restoration and livelihood development in Central
Asia. East Central Asia. World Agroforestry Centre.
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 47

Fanadzo, M., Dalicuba, M., & Dube, E. (2018). Application of conservation agriculture principles
for the management of field crops pests. In Sustainable agriculture reviews 28: Ecology for
agriculture (pp. 125–152). Springer.
Ferroni, M., & Castle, P. (2011). Public-private partnerships and sustainable agricultural develop-
ment. Sustainability, 3(7), 1064–1073. https://doi.org/10.3390/su3071064
Flanagan, K., Robertson, K., & Hanson, C. (2019). Reducing food loss and waste. Setting the
global action agenda. World Resources Institute (WRI).
Furumo, P. R., & Lambin, E. F. (2020). Scaling up zero-deforestation initiatives through public-­
private partnerships: A look inside post-conflict Colombia. Global Environmental Change, 62,
102055. https://doi.org/10.1016/j.gloenvcha.2020.102055
Ghosh, S., Sarkar, T., Pati, S., Kari, Z. A., Edinur, H. A., & Chakraborty, R. (2022). Novel bioac-
tive compounds from marine sources as a tool for functional food development. Frontiers in
Marine Science, 9, 832957. https://doi.org/10.3389/fmars.2022.832957
Giller, K. E., Beare, M. H., Lavelle, P., Izac, A. M., & Swift, M. J. (1997). Agricultural intensifica-
tion, soil biodiversity and agroecosystem function. Applied Soil Ecology, 6(1), 3–16. https://
doi.org/10.1016/S0929-­1393(96)00149-­7
Gunders, D., & Bloom, J. (2017). Wasted: How America is losing up to 40 percent of its food from
farm to fork to landfill. Natural Resources Defense Council.
Gupta, N., Pradhan, S., Jain, A., & Patel, N. (2021). Sustainable agriculture in India
2021. CEEW Report, 122p. https://www.ceew.in/sites/default/files/CEEWFOLU-­
Sustainable-­Agriculture-­in-­India-­2021-­20Apr21
Hermans, F., Geerling-Eiff, F., Potters, J., & Klerkx, L. (2019). Public-private partnerships as
systemic agricultural innovation policy instruments–Assessing their contribution to innovation
system function dynamics. NJAS-Wageningen Journal of Life Sciences, 88, 76–95. https://doi.
org/10.1016/j.njas.2018.10.001
Heydari, M., Lai, K. K., & Xiaohu, Z. (2020). Risk management in public-private partnerships.
Routledge.
Kampragou, E., Lekkas, D. F., & Assimacopoulos, D. (2011). Water demand management:
Implementation principles and indicative case studies. Water and Environment Journal, 25(4),
466–476. https://doi.org/10.1111/j.1747-­6593.2010.00240.x
Keers, B. B., & van Fenema, P. C. (2018). Managing risks in public-private partnership forma-
tion projects. International Journal of Project Management, 36(6), 861–875. https://doi.
org/10.1016/j.ijproman.2018.05.001
Khanal, S., Fulton, J., & Shearer, S. (2017). An overview of current and potential applications of
thermal remote sensing in precision agriculture. Computers and Electronics in Agriculture,
139, 22–32. https://doi.org/10.1016/j.compag.2017.05.001
Koul, B., Yakoob, M., & Shah, M. P. (2022). Agricultural waste management strategies for envi-
ronmental sustainability. Environmental Research, 206, 112285. https://doi.org/10.1016/j.
envres.2021.112285
Kumar, V., Wankhede, K. G., & Gena, H. C. (2015). Role of cooperatives in improving livelihood
of farmers on sustainable basis. American Journal of Educational Research, 3(10), 1258–1266.
Kumawat, T. K., Sharma, V., Kumawat, V., Pandit, A., & Biyani, M. (2022). Agricultural and
agro-wastes as sorbents for remediation of noxious pollutants from water and wastewater.
In Sustainable materials for sensing and remediation of noxious pollutants (pp. 161–176).
Elsevier. https://doi.org/10.1016/B978-­0-­323-­99425-­5.00017-­7
Kusumowardani, N., Tjahjono, B., Lazell, J., Bek, D., Theodorakopoulos, N., Andrikopoulos, P.,
& Priadi, C. R. (2022). A circular capability framework to address food waste and losses in the
agri-food supply chain: The antecedents, principles and outcomes of circular economy. Journal
of Business Research, 142, 17–31. https://doi.org/10.1016/j.jbusres.2021.12.020
Laghari, A. N., Vanham, D., & Rauch, W. (2012). The Indus basin in the framework of current and
future water resources management. Hydrology and Earth System Sciences, 16(4), 1063–1083.
https://doi.org/10.5194/hess-­16-­1063-­2012
48 P. Kumar et al.

Lal, R. (2015). Restoring soil quality to mitigate soil degradation. Sustainability, 7(5), 5875–5895.
https://doi.org/10.3390/su7055875
Lal, R., Delgado, J. A., Groffman, P. M., Millar, N., Dell, C., & Rotz, A. (2011). Management to
mitigate and adapt to climate change. Journal of Soil and Water Conservation, 66(4), 276–285.
https://doi.org/10.2489/jswc.66.4.276
Li, C., Mirosa, M., & Bremer, P. (2020). Review of online food delivery platforms and their
impacts on sustainability. Sustainability, 12(14), 5528. https://doi.org/10.3390/su12145528
Liang, Z., Liu, X., Xiong, J., & Xiao, J. (2020). Water allocation and integrative management
of precision irrigation: A systematic review. Water, 12(11), 3135. https://doi.org/10.3390/
w12113135
Liehr, S., Brenda, M., Cornel, P., Deffner, J., Felmeden, J., Jokisch, A., & Urban, W. (2016).
From the concept to the tap—Integrated water resources management in Northern Namibia. In
Integrated water resources management: Concept, research and implementation (pp. 683–717).
Springer.
Maleksaeidi, H., & Karami, E. (2013). Social-ecological resilience and sustainable agriculture
under water scarcity. Agroecology and Sustainable Food Systems, 37(3), 262–290. https://doi.
org/10.1080/10440046.2012.746767
Manda, R. R., Addanki, V. A., & Srivastava, S. (2021). Role of drip irrigation in plant health
management, its importance and maintenance. Plant Archives, 21(1), 1294–1302. https://doi.
org/10.51470/PLANTARCHIVES.2021.v21.S1.204
Mandari, V., & Devarai, S. K. (2021). Biodiesel production using homogeneous, heterogeneous,
and enzyme catalysts via transesterification and esterification reactions: A critical review.
Bioenergy Research, 1–27. https://doi.org/10.1007/s12155-­021-­10333-­w
Meena, B. L., Fagodiya, R. K., Prajapat, K., Dotaniya, M. L., Kaledhonkar, M. J., Sharma, P. C.,
et al. (2018). Legume green manuring: An option for soil sustainability. In Legumes for soil
health and sustainable management (pp. 387–408). Springer.
Nawaz, A., Sufyan, M., Gogi, M. D., & Javed, M. W. (2019). Sustainable management of insect-­
pests. In Innovations in sustainable agriculture (pp. 287–335). Springer.
Newman, P., & Jennings, I. (2012). Cities as sustainable ecosystems: Principles and practices.
Island Press.
Osmond, D., Meals, D., Hoag, D., Arabi, M., Luloff, A., Jennings, G., & Line, D. (2012).
Improving conservation practices programming to protect water quality in agricultural water-
sheds: Lessons learned from the National Institute of Food and Agriculture–Conservation
Effects Assessment Project. Journal of Soil and Water Conservation, 67(5), 122A–127A.
Piñeiro, V., Arias, J., Dürr, J., Elverdin, P., Ibáñez, A. M., Kinengyere, A., & Torero, M. (2020).
A scoping review on incentives for adoption of sustainable agricultural practices and their
outcomes. Nature Sustainability, 3(10), 809–820. https://doi.org/10.1038/s41893-­020-­00617-­y
Pingali, P., Khwaja, Y., & Meijer, M. (2005). Commercializing small farms: Reducing transaction
cost. http://www.fao.org/3/a-­af144t.pdf
Rahman, M. M., Khan, I., Field, D. L., Techato, K., & Alameh, K. (2022). Powering agriculture:
Present status, future potential, and challenges of renewable energy applications. Renewable
Energy, 188, 731–749. https://doi.org/10.1016/j.renene.2022.02.065
Reddy, A. (2018). Impact study of Paramparagath Krishi Vikas Yojana (Organic Agriculture)
Scheme of India. Reddy A Amarender (2018) Impact study of Paramparagat Krishi Vikas Yojana,
National Institute of Agricultural Extension Management (MANAGE), Hyderabad-500030,
p. 210. https://doi.org/10.2139/ssrn.3249954
Rohini, C., Geetha, P. S., Vijayalakshmi, R., Mini, M. L., & Pasupathi, E. (2020). Global effects of
food waste. Journal of Pharmacognosy and Phytochemistry, 9(2), 690–699.
Roy, T., & George, K. J. (2020). Precision farming: A step towards sustainable, climate-smart
agriculture. In Global climate change: Resilient and smart agriculture (pp. 199–220). Springer.
Russo, T., Alfredo, K., & Fisher, J. (2014). Sustainable water management in urban, agricultural,
and natural systems. Water, 6(12), 3934–3956. https://doi.org/10.3390/w6123934
Approach to Reduce Agricultural Waste via Sustainable Agricultural Practices 49

Searchinger, T. D., Malins, C., Dumas, P., Baldock, D., Glauber, J., Jayne, T., & Marenya, P. (2020).
Revising public agricultural support to mitigate climate change. World Bank Publications.
Sen, S., Ganguly, S., Das, A., Sen, J., & Dey, S. (2016). Renewable energy scenario in India:
Opportunities and challenges. Journal of African Earth Sciences, 122, 25–31. https://doi.
org/10.1016/j.jafrearsci.2015.06.002
Shah, F., & Wu, W. (2019). Soil and crop management strategies to ensure higher crop produc-
tivity within sustainable environments. Sustainability, 11(5), 1485. https://doi.org/10.3390/
su11051485
Sharma, R., & Bhardwaj, S. (2017). Effect of mulching on soil and water conservation-A review.
Agricultural Reviews, 38(4), 311–315.
Sindhu, R., Gnansounou, E., Binod, P., & Pandey, A. (2016). Bioconversion of sugarcane crop
residue for value added products–An overview. Renewable Energy, 98, 203–215. https://doi.
org/10.1016/j.renene.2016.02.057
Singh, M. (2021). Organic farming for sustainable agriculture. Indian Journal of Organic Farming,
1(1), 1–8.
Sivanappan, R. K. (2006, November). Rainwater harvesting, conservation and management strate-
gies for urban and rural sectors. In National seminar on rainwater harvesting and water man-
agement (Vol. 11, No. 12, p. 1). Institution of Engineers (India), Nagpur Local Centre, Nagpur
in Association with UNESCO.
Smith, P., House, J. I., Bustamante, M., Sobocká, J., Harper, R., Pan, G., & Pugh, T. A. (2016).
Global change pressures on soils from land use and management. Global Change Biology,
22(3), 1008–1028. https://doi.org/10.1111/gcb.13068
Smyth, S. J., Webb, S. R., & Phillips, P. W. (2021). The role of public-private partnerships in
improving global food security. Global Food Security, 31, 100588. https://doi.org/10.1016/j.
gfs.2021.100588
Soltani, M., Kashkooli, F. M., Souri, M., Rafiei, B., Jabarifar, M., Gharali, K., & Nathwani,
J. S. (2021). Environmental, economic, and social impacts of geothermal energy sys-
tems. Renewable and Sustainable Energy Reviews, 140, 110750. https://doi.org/10.1016/j.
rser.2021.110750
Somasundaram, J., Sinha, N. K., Dalal, R. C., Lal, R., Mohanty, M., Naorem, A. K., & Chaudhari,
S. K. (2020). No-till farming and conservation agriculture in South Asia–Issues, challenges,
prospects and benefits. Critical Reviews in Plant Sciences, 39(3), 236–279. https://doi.org/1
0.1080/07352689.2020.1782069
Soundarrajan, P., & Vivek, N. (2016). Green finance for sustainable green economic growth in India.
Agricultural Economics, 62(1), 35–44. https://doi.org/10.17221/174/2014-­AGRICECON
Syme, G. J., Reddy, V. R., & Ranjan, R. (2015). Justice and equity in watershed development
in Andhra Pradesh. In Integrated assessment of scale impacts of watershed intervention
(pp. 317–352). Elsevier. https://doi.org/10.1016/B978-­0-­12-­800067-­0.00010-­4
Taylor, B. M., & Van Grieken, M. (2015). Local institutions and farmer participation in agri-­
environmental schemes. Journal of Rural Studies, 37, 10–19. https://doi.org/10.1016/j.
jrurstud.2014.11.011
Umar, B. B., Aune, J. B., Johnsen, F. H., & Lungu, O. I. (2011). Options for improving smallholder
conservation agriculture in Zambia. Journal of Agricultural Science, 3(3), 50.
Umesha, S., Manukumar, H. M., & Chandrasekhar, B. (2018). Sustainable agriculture and food
security. In Biotechnology for sustainable agriculture (pp. 67–92). Woodhead Publishing.
https://doi.org/10.1016/B978-­0-­12-­812160-­3.00003-­9
Vera, I., Wicke, B., Lamers, P., Cowie, A., Repo, A., Heukels, B., & van der Hilst, F. (2022). Land
use for bioenergy: Synergies and trade-offs between sustainable development goals. Renewable
and Sustainable Energy Reviews, 161, 112409. https://doi.org/10.1016/j.rser.2022.112409
Wato, T., Amare, M., Bonga, E., Demand, B. B. O., & Coalition, B. B. R. (2020). The agricultural
water pollution and its minimization strategies–A review. Journal of Resources Development
and Management, 64, 10–22.
50 P. Kumar et al.

Wirza, R., & Nazir, S. (2021). Urban aquaponics farming and cities-A systematic literature review.
Reviews on Environmental Health, 36(1), 47–61. https://doi.org/10.1515/reveh-­2020-­0064
Yadav, M., Vashisht, B. B., Jalota, S. K., Kumar, A., & Kumar, D. (2022). Sustainable water man-
agement practices for intensified agriculture. In Soil-water, agriculture, and climate change:
Exploring linkages (pp. 131–161). Springer International Publishing.
Yang, Y., Tilman, D., Furey, G., & Lehman, C. (2019). Soil carbon sequestration accelerated
by restoration of grassland biodiversity. Nature Communications, 10(1), 718. https://doi.
org/10.1038/s41467-­019-­08636-­w
Young, C. W., Russell, S. V., Robinson, C. A., & Chintakayala, P. K. (2018). Sustainable retail-
ing–influencing consumer behaviour on food waste. Business Strategy and the Environment,
27(1), 1–15. https://doi.org/10.1002/bse.1966
Biomass Waste and Bioenergy Production:
Challenges and Alternatives

Ahmed Albahnasawi , Murat Eyvaz , Motasem Y. D. Alazaiza ,


Nurullah Özdoğan , Ercan Gurbulak , Sahar Alhout ,
and Ebubekir Yuksel

1 Exploring the Newest Developments in Biowaste


Conversion Technologies and Current Trends
in Bioresource and Waste Management

The attention given to biomass derived from biowastes is highly significant due to
their exceptional physical, chemical, and biological properties. These remarkable
characteristics make them exceptionally well suited for a wide range of applications
within the realm of biorefinery. They serve as valuable precursors for biofuels,
chemicals, and various other biomaterials, contributing to the sustainable use of
resources (Guan et al., 2022). Moreover, these waste materials offer a promising
avenue as alternative and renewable energy sources in diverse forms (Fig. 1).
Recognizing the limitations and detrimental effects associated with current waste
management approaches, it becomes imperative to focus on the development of
sustainable waste treatment technologies. These innovative technologies offer sev-
eral advantages. They encompass waste-activated sludge pretreatment, pyrolytic
processes that enable biochar production, biohydrogen production derived from lig-
nocellulosic biomass, anaerobic fermentation utilizing herbal residues, algal bio-
mass production, syngas production, biomethanation, enzymatic hydrolysis for

A. Albahnasawi (*) · M. Eyvaz · E. Gurbulak · E. Yuksel


Department of Environmental Engineering, Gebze Technical University, Kocaeli, Turkey
M. Y. D. Alazaiza
Department of Civil and Environmental Engineering, College of Engineering,
A’Sharqiyah University, Ibra, Oman
N. Özdoğan
Department of Environmental Engineering, Bursa Uludag University, Bursa, Turkey
S. Alhout
Department of Pharmacy and Biotechnology, University of Palestine, Al-Zahra, Palestine

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 51


A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_3
52 A. Albahnasawi et al.

Fig. 1 Biowaste sources

biorefinery applications, and bioconversion of food waste via dark fermentation


(Ladakis et al., 2022). By employing these techniques, successful bioprocesses can
be achieved, simultaneously addressing the challenge of efficient bioresource man-
agement across agro-industrial, food waste, and algal biomass sectors (Saratale
et al., 2022).
In a study conducted by Suriapparao et al. (2022), they delved into the interac-
tion and pyrolysis mechanism between plastic wastes and agro-residuals as feed-
stocks. Their findings revealed a remarkably high pyrolysis conversion index
(99–100%), indicating that the proposed mechanism played a pivotal role in the
sustainable management of biowaste containing hemicellulose, lignin, and cellu-
lose, particularly for energy-related purposes. According to Pal et al. (2022), the
co-digestion of rice straw and cow dung can significantly enhance the concentration
of biomethane, reaching as high as 58.30%, while maintaining low levels of H2S. As
an alternative to green waste, food waste can undergo a direct two-stage anaerobic
digestion process involving primary lactate-type fermentation. This approach allows
for a doubled bioenergy recovery rate compared with a one-stage process. In another
study by Tahir et al. (2022), lignocellulosic biomass, specifically Paulownia waste,
was explored for green hydrogen production. The addition of SnO2 nano-catalysts
was found to increase the biohydrogen yield by approximately 47%. Kim et al.
(2022) reported that a conductive metal compound can serve as a catalyst to enhance
biohydrogen production through dark fermentation. The supplementation of mag-
netite resulted in a 25.60% increase in biohydrogen production. However, a more
promising alternative for conventional green hydrogen production involves a com-
bination of ultrasonic-assisted alkaline pretreatment, dark fermentation, and
Biomass Waste and Bioenergy Production: Challenges and Alternatives 53

microbial electrolysis cells. This integrated approach was shown using hyacinth
water as an example, resulting in a bio-H2 yield approximately four times higher
than that achieved with only dark fermentation/microbial electrolysis cell, as high-
lighted by Thu Ha Tran and Khanh Thinh Nguyen (2022)
According to the findings of Liang et al. (2022), a stable bacterial community
consisting of Prevotella, Rikenellaceae_RC9_gut_group, Ruminococcus, and
Succiniclasticum demonstrated remarkable efficiency in hydrolyzing hemicellulose
and cellulose. The hydrolysis effectiveness ranged from 36.50% to 52.20% for
hemicellulose and 29.40% to 38.40% for cellulose. In a separate study conducted by
Zhu et al. (2022), the application of p-toluenesulfonic acid combined with hydrogen
peroxide-assisted pretreatment proved to be highly beneficial for enhancing the pro-
duction of fermentable sugars from walnut shells. This innovative approach resulted
in a simultaneous increase in glucose yield of up to 94.40%. Behera et al. (2022)
undertook a comprehensive review of microalgae biomass as a promising alterna-
tive source of bioenergy. They specifically explored hydrochar production through
the hydrothermal carbonization (HTC) process. Additionally, they emphasized the
potential use of Microcystis sp., an algal biomass, in the eco-friendly production of
fertilizers. By treating the produced hydrochar with 1% citric acid, they achieved an
impressive recovery rate of 95% for phosphorus (P) and a 34% increase in nutrient
use efficiency compared with other chemical alternatives. In the realm of food waste
digestate composting as a climate-friendly biofertilizer, Li et al. (2022) determined
that the optimal dosage of zeolite addition was 10%. This specific dosage not only
substantially improved the degradation rate by approximately 57%, but also led to
significant reductions of carbon loss by about 43.1% and nitrogen loss by about
5.68%. Moreover, it effectively mitigated emissions of NH3 and N2O by approxi-
mately 45%.
The transition from traditional fossil fuels like gasoline and natural gas to renew-
able alternatives poses a significant challenge. Biowastes and green wastes emerge
as highly promising candidates due to the projected growth of the biofuel and
renewable energy sectors. The annual growth rate is expected to reach at least
1.20%, eventually amounting to approximately 2.80% by 2035 (Piechota
et al., 2023).

2 Approach to Circular Economy (CE) Challenges


and Conclusions

The escalating strain on our finite natural resources has captured global attention,
prompting the implementation of strategic plans and initiatives for sustainable
development. In this context, a recent study has delved into the potential of convert-
ing discarded coffee grounds into a diverse array of valuable bioproducts, encom-
passing biodiesel, bioethanol, bio-ether, bio-oil, biochar, biogas, and green
biocomposites, alongside other high-value products with lower yields (Leong &
Chang, 2022).
54 A. Albahnasawi et al.

Fig. 2 Biowaste composting for circular economy approach for biowaste

The existing body of literature primarily concentrates on the economic dimen-


sions of the circular economy (Fig. 2), often overlooking the ecological and social
perspectives that are equally vital. Conducting surveys becomes crucial to identify
the industrial challenges that appear during the transition from a linear to a circular
economy, enabling the search for practical solutions to overcome these obstacles. It
is imperative to consider the implications of consumers embracing circular econ-
omy practices during the end-use phase to ensure the establishment of a sustainable
circular bioeconomy. Notably, J. K. Saini et al. (2022) shed light on the potential
hurdles encountered in the enzymatic bioconversion of lignocellulosic materials
into biofuels and value-added chemicals within the framework of a biorefinery.
Technical challenges arise, particularly in the realm of bio-based energy produc-
tion, where the choice of biomass plays a pivotal role. Addressing financial chal-
lenges involves the implementation of biorefineries to effectively manage residual
materials. Social challenges need improved communication and awareness sur-
rounding biorefineries and their associated products to enhance their demand and
market value. Consequently, the development of advanced skills and modifications
specific to the feedstocks and expected bioproducts, coupled with government assis-
tance, becomes crucial in ensuring the security of capital investments and mitigating
adverse environmental impacts. Given the importance of these factors, undertaking
meticulous and comprehensive biotechnological investigations becomes necessary
to optimize financial returns while minimizing energy consumption and overall
investment, thereby propelling the circular bioeconomy forward. This approach
serves as a blueprint for the synergistic conversion of agricultural waste into
Biomass Waste and Bioenergy Production: Challenges and Alternatives 55

value-added products, facilitating the efficient circulation of essential nutrients.


This succinct assessment aims to underscore the substantial potential inherent in
processing biowaste sources to obtain technically, nutritionally, and organically
valuable molecules, thereby inspiring the scientific and engineering communities to
explore novel and more efficient waste conversion methods. Despite the vast poten-
tial and availability of biowastes, their proper use is still largely unrealized.
Moreover, the study highlights how advanced bioconversion methods can supplant
traditional composting techniques, yielding high-value compost and nonconven-
tional energy. Furthermore, it offers valuable insights into the research papers pub-
lished in the virtual special issue (VSI) of the Bioresource Technology Journal. The
findings and research presented in this article significantly contribute to sustainable
development and energy efficiency in harnessing the potential of biowastes. Finally,
it underscores the necessity for well-organized and innovative research schemes to
effectively address the accumulation of biowastes.

3 Energy Crisis and Biowaste Energy Potential

Modern society’s heavy dependence on fossil fuels as the primary energy source is
well-documented (Chuah et al., 2021). However, the Intergovernmental Panel on
Climate Change has underscored the urgency of the situation, emphasizing the
imperative need to transition away from fossil fuels toward cleaner energy sources
in order to mitigate the devastating effects of climate change on humanity (Jain
et al., 2022). This critical context has spurred researchers to explore innovative tech-
nologies and processes that assess the environmental implications of different prod-
ucts (Chuah et al., 2022). A wealth of studies has shown the significant potential for
generating bioenergy through the efficient utilization of biowaste (Bokhari et al.,
2019). Figure 3 shows the value-added products derived from biowaste.
The rapid pace of urbanization, industrialization, population growth, and the
rise in consumerism have resulted in a substantial global production of biomass
waste (Khan et al., 2020). Remarkably, approximately 44% (w/w) of all biowastes
consist of solid waste, showing diverse characteristics and compositions.
Astonishingly, only a fraction of the total municipal waste generated by 35 mem-
ber countries of the Organization for Economic Cooperation and Development,
accounting for 44% (w/w), undergoes bio-based processing techniques to produce
economically valuable compounds (Vakalis et al., 2017). Industries, the agricul-
tural sector, and households regularly generate various biowastes that hold
immense potential as sources for bioenergy production (Escamilla-Alvarado
et al., 2017). In 2016 alone, a staggering 2.1 × 104 metric tons of biowaste was
generated globally, and this figure is projected to rise to 2.2 × 104 metric tons by
2025. If sustainable waste management strategies are not effectively implemented,
it is estimated that by 2050, a staggering 3.4 × 104 metric tons of biowaste could
be generated (Paes et al., 2019). This pressing need underscores the importance of
56 A. Albahnasawi et al.

Fig. 3 Conversion biowaste to value-added products (Xu et al., 2022)

transforming biowastes into valuable bioproducts and bioenergy, emphasizing the


urgency to address this challenge (Abdelghaffar, 2021).
The biowaste-to-energy (BtE) technologies hold immense potential, estimated to
generate approximately 26 billion US dollars, with expectations to reach 40 billion
dollars by 2023. However, the lack of effective mitigation strategies results in the
disposal of substantial quantities of biowaste in landfills, leading to the significant
production of methane gas during decomposition and posing alarming environmen-
tal consequences. In the United States, landfills rank as the third-largest source of
methane generation, contributing significantly to the overall temperature rise due to
the greenhouse effect (Sotiropoulos et al., 2016). At the core of the circular econ-
omy lies the principle of promoting the reuse and recycling of products. Developing
countries play a pivotal role in using most environmental resources while producing
goods that yield both environmental and economic benefits (Teigiserova et al.,
2020). Key aspects of the circular economy encompass conducting life cycle assess-
ments of products, designing strategies to optimize resource utilization, and effec-
tively managing biowaste (Awasthi et al., 2021). The global bioeconomy aims to
ensure the sustainable management of environmental resources by focusing on asset
Biomass Waste and Bioenergy Production: Challenges and Alternatives 57

viability and biomass sustainability (Wainaina et al., 2020). The concepts of the
bioeconomy and circular economy align harmoniously, as a circular bioeconomy
involves transforming waste products into clean energy and other valuable resources
through recycling and efficient management practices. Governments worldwide are
increasingly prioritizing the bioeconomy as a crucial part of sustainable develop-
ment, particularly in addressing the management of urban waste (Sodhi et al., 2022).
Several obstacles hinder the implementation of a circular economy, including the
severe environmental impact of landfilling, excessive dependence on heavy indus-
tries, and the rapid growth of urban populations (Jain et al., 2022). According to
Sodhi et al. (2022), a wide range of degradable biowaste, including food waste,
agricultural waste, kitchen waste, green waste, sewage, sludge, agro-industry, and
forestry residues, can be effectively managed through their conversion into bioprod-
ucts, composites (Cheng et al., 2020), nanomaterials (Cao et al., 2022), nanotube
sheets (Wang et al., 2021), and biofuel (Kumar Awasthi et al., 2022). Traditionally,
organic waste management has relied on landfilling, composting, and incineration,
each with its own advantages and disadvantages. However, contemporary society is
increasingly embracing bio-based processing technologies as alternatives to con-
ventional methods due to their lower energy requirements, reduced investment
costs, and higher recovery rates of value-added products, thereby aligning with the
principles of the circular economy (Rasapoor et al., 2020). Two key processes,
anaerobic digestion and microbial degradation, play crucial roles in improving
nutrient circulation from organic wastes (Zamri et al., 2021). Biorefineries provide
the capacity to convert organic wastes into energy and valuable products (Bokhari
et al., 2020). Processed wastes, both in liquid and solid forms, contain a wealth of
minerals, proteins, and carbohydrates that can be transformed into enzymes, bioac-
tive compounds, and pigments with diverse applications in therapeutics and indus-
try (Cheng et al., 2020). The rapid pace of urbanization and population growth
contributes to the significant generation of biowaste, originating from various
sources (Escamilla-Alvarado et al., 2017). These biowastes hold immense potential
for conversion into value-added products, such as bioenergy (Abdelghaffar, 2021).
Transformation techniques like microbial fuel cells (Raychaudhuri et al., 2021) and
bioreactors (Park et al., 2021) can be used to effectively harness the potential of
biowaste.

4 Feedstocks for Bioenergy Production

A significant portion of organic solid waste includes agricultural and food waste.
Conventional methods employed to manage these waste types include incineration,
composting, landfilling, and their utilization as animal feed (Kumar et al., 2022).
However, incineration reduces the recovery of valuable nutrients from the waste,
while landfilling presents issues such as uncontrolled microbial degradation and the
generation of greenhouse gases (Mehariya et al., 2021). Unfortunately, due to inad-
equate mitigation measures, substantial quantities of biowaste are currently being
58 A. Albahnasawi et al.

discarded in landfills, resulting in the significant production of methane gas during


the decomposition process. In fact, landfills rank as the third major source of meth-
ane generation in the United States (Sotiropoulos et al., 2016). Moreover, the
decomposition of organic matter in landfills leads to the release of hydrogen sulfide
(H2S) and organic mercaptans, giving rise to odor-related problems. The volatile
compounds emitted from landfills contribute to air pollution. Additionally, the
deposition of various industrial solid wastes in landfills can lead to the mixing of
surface water with the waste, thus contaminating groundwater and altering its qual-
ity. Waste treatment processes also transfer substances that directly affect the qual-
ity of soil, air, and water. Leachate, a liquid formed during the decomposition of
municipal solid waste, has organic components and heavy metals (Dastjerdi et al.,
2021). While using agricultural and food wastes as animal feed is cost-effective, it
needs regulated implementation due to the unknown composition of the waste.
Composting has appeared as a well-studied approach for managing these waste
types, as it generates biofertilizer that can be beneficial for farmers (Su et al., 2020).

5 Bioreactor Development for Energy Production

Bioreactors play a crucial role in facilitating cellular growth and metabolism by


providing an optimal environment. They are particularly important for microbial
energy conversion, which is essential for achieving sustainability in energy-­intensive
processes. The design of bioreactors holds significant importance in energy produc-
tion as it offers several advantages, including process simplicity, cost reduction of
raw materials, decreased carbon footprint, precise control over environmental
parameters, improved product yield, and efficient conversion of raw materials into
fuels (Xu et al., 2018). Considering the wide range of bioenergy applications, which
include both liquid fuels and gases, extensive research has led to the development of
various bioreactor designs. Previous studies have contributed to the advancement of
these bioreactors, catering to the specific requirements of producing different types
of biofuels as sustainable energy sources (Choudri & Baawain, 2016). These spe-
cialized bioreactors are designed to optimize the production of diverse biofuels,
contributing to the advancement and implementation of sustainable energy solutions.

5.1 Bioreactors for Biohydrogen Production

Biohydrogen is widely acknowledged as a sustainable alternative to chemically syn-


thesized hydrogen due to its high energy content and environmentally friendly con-
version process. Although biohydrogen production is still in its early stages, the
increasing demand has spurred significant advancements in biohydrogen produc-
tion using various types of bioreactors, ranging from small-scale laboratory systems
to large-scale commercial setups. The design of bioreactors is influenced by several
Biomass Waste and Bioenergy Production: Challenges and Alternatives 59

factors, including the type and concentration of the feedstock, temperature, partial
pressure of hydrogen, hydraulic retention time (HRT), and pH of the culture, as
these variables have a direct impact on biohydrogen production (Jabbari et al., 2019).
Various types of photobioreactors, such as pond or pool-type, tubular, and flat-­
plate reactors, have been employed for biohydrogen production using diatoms,
microalgae, and cyanobacteria. These microorganisms exhibit variations in their
photochemical efficiency and light absorption characteristics, which in turn influ-
ence the performance of photobioreactors (Park et al., 2021). Additionally, different
types of reactors, including batch, semicontinuous, and continuous stirred tank
reactors (CSTRs), have been utilized for biohydrogen production. Among these,
continuous reactors have demonstrated the highest productivity, while batch reac-
tors yield the lowest biohydrogen production rates (Garcia-Peña et al., 2018).
In addition to these reactor types, various other bioreactor designs have been
developed to enhance biohydrogen yields, such as membrane bioreactors (MBRs),
fluidized bed bioreactors, anaerobic sequencing batch reactors (ASBRs), anaerobic
sludge blanket (UASB) bioreactors, and fixed-bed bioreactors. More recently,
microbial electrolysis cells (MECs) that use electro-hydrogenesis have been
explored for biohydrogen production (Sim et al., 2021). However, the efficiency of
biohydrogen production in MECs with nominal electrical inputs is not yet satisfac-
tory, leading to the development of two-stage hybrid reactors. In this approach, the
first stage involves the dark fermentation of biomass into acetate, hydrogen, and
carbon dioxide, while the second stage focuses on converting acetate into hydrogen
and carbon dioxide (Lee et al., 2009). Hybrid systems offer significant advantages
over conventional methods in terms of both yield and volumetric production rates of
hydrogen. Despite the promising developments in biohydrogen production, achiev-
ing sustainable and rapid biohydrogen production soon needs the proper design and
optimization of bioreactors.

5.2 Bioreactors for Biodiesel Production

The depletion of petroleum reserves has spurred the development of alternative


renewable fuels like biodiesel, which plays a crucial role in meeting global energy
demands. The economic viability of biodiesel production relies heavily on the
selection of raw materials and the design of bioreactors. Initially, packed bed reac-
tors with limited capacity were employed, but they encountered challenges such as
insufficient maintenance of feed solution velocity above the minimum fluidization
velocity and clogging of finer particulate matter. To overcome these issues, fluidized
bed bioreactors were introduced, where the superficial velocity of the feed solution
exceeded the fluidization velocity. This advancement enabled higher-capacity oper-
ation and reduced operating costs. Additionally, the development of semi-fluidized
bed bioreactors combined the advantages of both packed bed and fluidized bed sys-
tems. These reactors featured a higher fluid velocity compared with conventional
60 A. Albahnasawi et al.

fluidized bed configurations and incorporated a porous plate to control bed expan-
sion during the process (da Costa et al., 2021).
Moreover, inverse fluidized bed bioreactors were introduced, with the feed solu-
tion introduced from the top and flowing down the column under the force of grav-
ity. These reactors maintained an operating superficial velocity 20–25% higher than
the minimum inverse fluidization velocity (Narayanan & Pandey, 2018). Recent
technological advancements have facilitated the mathematical assessment of biore-
actor performance using modeling and computer-aided design (CAD) for biodiesel
production (Ding et al., 2010). The accuracy of CAD-developed architectures has
been validated through experimental data, leading to successful industrial-scale bio-
diesel production (Abomohra et al., 2018). While significant progress has been
made in bioreactor development for biodiesel production, further extensive studies
are needed to fully commercialize this technology on a global scale.

5.3 Bioreactors for Biogas Production

Biogas, a renewable and sustainable alternative to fossil fuels, holds great promise
for meeting the global energy demand by using various biomass sources. It consists
of gases such as methane, carbon dioxide, and hydrogen, which are produced
through the anaerobic breakdown of organic matter (Zhao et al., 2021). These gases
can be further utilized through oxidation or combustion in the presence of oxygen,
yielding an energy potential of approximately 28.8 MJ/MJ. The production of bio-
gas relies on two key operating parameters: the type of organic matter and the choice
of reactor design. Extensive research has been conducted on different reactor con-
figurations, including anaerobic contact reactors, continuous stirred tank reactors
(CSTRs), fixed film reactors (FFRs), fluidized bed reactors, up-flow anaerobic
sludge blanket reactors, expanded granular sludge bed reactors, and jet flow anaero-
bic bioreactors (Kapoor et al., 2019). Among these options, CSTRs and FFRs are
the most used. However, CSTRs suffer from limited bacterial retention, while FFRs
offer more efficient degradation of complex organic matter compared with CSTRs
(Mishra et al., 2021).
Recent advancements in computational fluid dynamics (CFD) have been instru-
mental in enhancing biogas production by reducing power consumption, improving
mixing performance, and gaining insights into flow characteristics affected by total
solid (TS) concentrations in organic waste matter (Saini et al., 2021). Anaerobic
ammonium oxidation (Annamox) bioreactors have attracted attention for their effi-
cient biogas production, effective nitrogen management, compact footprint, and
reduced costs (Wang et al., 2019). Nevertheless, further research is needed to com-
mercialize this technology and address technical challenges such as product inhibi-
tion, methane recovery, membrane fouling, and high costs.
The development of membrane bioreactors plays a critical role in biogas produc-
tion. In a recent study, the design and operation of an anaerobic bioreactor for
Biomass Waste and Bioenergy Production: Challenges and Alternatives 61

biogas production using wastewater were explored, with a specific focus on mem-
brane coupling and process fundamentals (Elmoutez et al., 2023). The study exam-
ined the performance of the anaerobic bioreactor at laboratory, pilot scale, and
prototype levels, investigating wastewater feeding, control strategies, and system
deficiencies. The combination of anaerobic ammonium oxidation and anaerobic
membrane bioreactors proved valuable in terms of resource recovery and energy
balance. Additionally, the study aimed to deepen our understanding of the biologi-
cal and physical aspects of the process (Elmoutez et al., 2023). Another study con-
centrated on the co-treatment of kitchen wastewater and food waste using a
two-stage anaerobic membrane bioreactor, showing improved process performance
during methanogenesis through acetogenesis and hydrolysis (Le et al., 2022).

5.4 Bioreactors for Bioethanol Production

The emergence of the bioethanol industry is a groundbreaking development in


achieving sustainable energy. With advancements in technology, there has been sig-
nificant progress in optimizing the fermentation process for bioethanol production,
leading to the use of various bioreactor configurations (Mahboubi et al., 2020).
Current research is focused on harnessing nonfood, cost-effective lignocellulosic
biomass for bioethanol production, which involves a multistep process encompass-
ing delignification, saccharification, and fermentation (Cremonez et al., 2021).
Consolidated bioprocessing has also been developed, utilizing improved strains that
enable simultaneous saccharification and fermentation. Bioethanol, as one of the
oldest bioenergy products, has driven the development of diverse bioreactor types,
including batch reactors, continuous stirred tank reactors (CSTRs), and various
membrane bioreactors, all of which have played a pivotal role in establishing biore-
fineries worldwide.
For example, a specifically designed bubble column reactor demonstrated suc-
cessful saccharification and bioethanol production through hydrothermal pretreat-
ment of lignocellulosic biomass. This method highlights its potential for application
in second-generation biorefineries. In another study, the operating conditions of a
microbial bioethanol production process were optimized, resulting in the develop-
ment of an optimal fed-batch bioreactor. The study validated a kinetic model for
bioethanol production at the laboratory scale, with a particular focus on the eco-
nomic and optimization aspects of bioreactor parameters. It proposed a model based
on a multiobjective dynamic optimization approach to systematically derive operat-
ing policies (Flevaris & Chatzidoukas, 2021). Furthermore, another investigation
explored syngas fermentation for bioethanol production using a tar-free bioreactor.
Charcoal and syngas were employed as substrates for Clostridium butyricum, and
the treatment of syngas resulted in higher colony-forming units (CFUs) compared
with untreated syngas. Lignocellulosic biomass served as the substrate for syngas
production in this study (Monir et al., 2020).
62 A. Albahnasawi et al.

6 Conclusion

Biowaste, once overlooked, has now emerged as a valuable feedstock capable of


producing a myriad of value-added products. Among the various applications aris-
ing from the processing of biowaste, bioenergy generation stands out as a particu-
larly promising avenue. The generation of bioenergy from biowaste necessitates a
thorough analysis and optimization of several parameters, including biowaste com-
position and conversion potential. The pursuit of innovative and sustainable tech-
nologies is crucial to effectively harness the vast potential of biowaste for enhanced
bioenergy production. Notably, feedstocks such as food, agriculture, beverage, and
municipal solid waste have proven to be viable resources for producing renewable
energy. However, despite its immense potential, the field of bioenergy generation
from biowaste is still in its early stages and requires more interdisciplinary research
to become a fully sustainable alternative to conventional energy sources. This chap-
ter presented here has systematically analyzed the bioconversion potential of bio-
waste into renewable energy. Furthermore, this chapter has addressed the importance
of bioreactor development for energy production, while also highlighting the major
challenges and future prospects in bioenergy production from biowaste. The com-
prehensive exploration of these topics underscores the potential of bioenergy pro-
duction from biowaste as a true game changer in waste valorization and energy
research. To realize the full potential of this excellent energy generation process, it
is imperative to approach its development in a systematic and strategic manner,
considering the techno-economic feasibilities. By doing so, bioenergy production
from biowaste can genuinely appear as a sustainable and viable alternative to con-
ventional energy sources, paving the way toward a greener and more sustain-
able future.

Declaration of Competing Interest The authors declare that they have no known competing
financial interests or personal relationships that could have appeared to influence the work reported
in this chapter.

References

Abdelghaffar, F. (2021). Biosorption of anionic dye using nanocomposite derived from chito-
san and silver nanoparticles synthesized via cellulosic banana peel bio-waste. Environmental
Technology & Innovation, 24, 101852. https://doi.org/10.1016/J.ETI.2021.101852
Abomohra, A. E. F., El-Naggar, A. H., & Baeshen, A. A. (2018). Potential of macroalgae for bio-
diesel production: Screening and evaluation studies. Journal of Bioscience and Bioengineering,
125(2), 231–237. https://doi.org/10.1016/J.JBIOSC.2017.08.020
Awasthi, M. K., Sarsaiya, S., Wainaina, S., Rajendran, K., Awasthi, S. K., Liu, T., Duan, Y., Jain, A.,
Sindhu, R., Binod, P., Pandey, A., Zhang, Z., & Taherzadeh, M. J. (2021). Techno-economics
and life-cycle assessment of biological and thermochemical treatment of bio-waste. Renewable
and Sustainable Energy Reviews, 144, 110837. https://doi.org/10.1016/J.RSER.2021.110837
Biomass Waste and Bioenergy Production: Challenges and Alternatives 63

Behera, B., Mari Selvam, S., & Balasubramanian, P. (2022). Hydrothermal processing of micro-
algal biomass: Circular bio-economy perspectives for addressing food-water-energy nexus.
Bioresource Technology, 359, 127443. https://doi.org/10.1016/J.BIORTECH.2022.127443
Bokhari, A., Chuah, L. F., Michelle, L. Z. Y., Asif, S., Shahbaz, M., Akbar, M. M., Inayat, A., Jami,
F., Naqvi, S. R., & Yusup, S. (2019). Microwave enhanced catalytic conversion of canola-­
based methyl ester: Optimization and parametric study. In Advanced biofuels: Applications,
technologies and environmental sustainability (pp. 153–166). https://doi.org/10.1016/B978-­0
-­08-­102791-­2.00006-­4
Bokhari, A., Yusup, S., Asif, S., Chuah, L. F., & Michelle, L. Z. Y. (2020). Process intensification
for the production of canola-based methyl ester via ultrasonic batch reactor: Optimization and
kinetic study. In Bioreactors: Sustainable design and industrial applications in mitigation of
GHG emissions (pp. 27–42). https://doi.org/10.1016/B978-­0-­12-­821264-­6.00003-­6
Cao, D., Malakooti, S., Kulkarni, V. N., Ren, Y., Liu, Y., Nie, X., Qian, D., Griffith, D. T., & Lu,
H. (2022). The effect of resin uptake on the flexural properties of compression molded sand-
wich composites. Wind Energy, 25(1), 71–93. https://doi.org/10.1002/WE.2661
Cheng, S. Y., Tan, X., Show, P. L., Rambabu, K., Banat, F., Veeramuthu, A., Lau, B. F., Ng, E. P.,
& Ling, T. C. (2020). Incorporating biowaste into circular bioeconomy: A critical review of
current trend and scaling up feasibility. Environmental Technology & Innovation, 19, 101034.
https://doi.org/10.1016/J.ETI.2020.101034
Choudri, B. S., & Baawain, M. (2016). Bioenergy from biofuel residues and
wastes. Water Environment Research, 88(10), 1446–1466. https://doi.org/10.217
5/106143016X14696400495217
Chuah, L. F., Klemeš, J. J., Bokhari, A., & Asif, S. (2021). A review of biodiesel production from
renewable resources: Chemical reactions. Chemical Engineering Transactions, 88, 943–948.
https://doi.org/10.3303/CET2188157
Chuah, L. F., Klemeš, J. J., Bokhari, A., Asif, S., Cheng, Y. W., Chong, C. C., & Show, P. L. (2022).
A review of intensification technologies for biodiesel production. In Biofuels and biorefining:
Volume 2: Intensification processes and biorefineries (pp. 87–116). https://doi.org/10.1016/
B978-­0-­12-­824117-­2.00009-­0
Cremonez, P. A., Teleken, J. G., Weiser Meier, T. R., & Alves, H. J. (2021). Two-stage anaerobic
digestion in agroindustrial waste treatment: A review. Journal of Environmental Management,
281, 111854. https://doi.org/10.1016/J.JENVMAN.2020.111854
da Costa, T. B., Simões, A. N., de Menezes, C. A., & Silva, E. L. (2021). Anaerobic biodegrada-
tion of biodiesel industry wastewater in mesophilic and thermophilic fluidized bed reactors:
Enhancing treatment and methane recovery. Applied Biochemistry and Biotechnology, 193(10),
3336–3350. https://doi.org/10.1007/S12010-­021-­03606-­9/TABLES/2
Dastjerdi, B., Strezov, V., Rajaeifar, M. A., Kumar, R., & Behnia, M. (2021). A systematic review
on life cycle assessment of different waste to energy valorization technologies. Journal of
Cleaner Production, 290, 125747. https://doi.org/10.1016/J.JCLEPRO.2020.125747
Ding, J., Wang, X., Zhou, X. F., Ren, N. Q., & Guo, W. Q. (2010). CFD optimization of continu-
ous stirred-tank (CSTR) reactor for biohydrogen production. Bioresource Technology, 101(18),
7005–7013. https://doi.org/10.1016/J.BIORTECH.2010.03.146
Elmoutez, S., Abushaban, A., Necibi, M. C., Sillanpää, M., Liu, J., Dhiba, D., Chehbouni, A., &
Taky, M. (2023). Design and operational aspects of anaerobic membrane bioreactor for effi-
cient wastewater treatment and biogas production. Environmental Challenges, 10, 100671.
https://doi.org/10.1016/J.ENVC.2022.100671
Escamilla-Alvarado, C., Poggi-Varaldo, H. M., & Ponce-Noyola, M. T. (2017). Bioenergy and
bioproducts from municipal organic waste as alternative to landfilling: A comparative life
cycle assessment with prospective application to Mexico. Environmental Science and Pollution
Research, 24(33), 25602–25617. https://doi.org/10.1007/S11356-­016-­6939-­Z/FIGURES/6
Flevaris, K., & Chatzidoukas, C. (2021). Optimal fed-batch bioreactor operating strategies for the
microbial production of lignocellulosic bioethanol and exploration of their economic impli-
cations: A step forward towards sustainability and commercialization. Journal of Cleaner
Production, 295, 126384. https://doi.org/10.1016/J.JCLEPRO.2021.126384
64 A. Albahnasawi et al.

Garcia-Peña, E. I., Niño-Navarro, C., Chairez, I., Torres-Bustillos, L., Ramírez-Muñoz, J., &
Salgado-Manjarrez, E. (2018). Performance intensification of a stirred bioreactor for fer-
mentative biohydrogen production. Preparative Biochemistry and Biotechnology, 48(1),
64–74. https://doi.org/10.1080/10826068.2017.1405269/SUPPL_FILE/LPBB_A_1405269_
SM4129.DOCX
Guan, R., Yuan, H., Yuan, S., Yan, B., Zuo, X., Chen, X., & Li, X. (2022). Current development
and perspectives of anaerobic bioconversion of crop stalks to biogas: A review. Bioresource
Technology, 349, 126615. https://doi.org/10.1016/J.BIORTECH.2021.126615
Jabbari, B., Jalilnejad, E., Ghasemzadeh, K., & Iulianelli, A. (2019). Recent progresses in applica-
tion of membrane bioreactors in production of biohydrogen. Membranes, 9(8), 100. https://doi.
org/10.3390/MEMBRANES9080100
Jain, A., Sarsaiya, S., Kumar Awasthi, M., Singh, R., Rajput, R., Mishra, U. C., Chen, J., & Shi,
J. (2022). Bioenergy and bio-products from bio-waste and its associated modern circular econ-
omy: Current research trends, challenges, and future outlooks. Fuel, 307, 121859. https://doi.
org/10.1016/J.FUEL.2021.121859
Kapoor, R., Ghosh, P., Kumar, M., & Vijay, V. K. (2019). Evaluation of biogas upgrading technolo-
gies and future perspectives: A review. Environmental Science and Pollution Research, 26(12),
11631–11661. https://doi.org/10.1007/S11356-­019-­04767-­1/FIGURES/3
Khan, M. B., Cui, X., Jilani, G., Tang, L., Lu, M., Cao, X., Sahito, Z. A., Hamid, Y., Hussain, B.,
Yang, X., & He, Z. (2020). New insight into the impact of biochar during vermi-stabilization
of divergent biowastes: Literature synthesis and research pursuits. Chemosphere, 238, 124679.
https://doi.org/10.1016/J.CHEMOSPHERE.2019.124679
Kim, D. H., Yoon, J. J., Kim, S. H., & Park, J. H. (2022). Acceleration of lactate-utilizing pathway
for enhancing biohydrogen production by magnetite supplementation in Clostridium butyricum.
Bioresource Technology, 359, 127448. https://doi.org/10.1016/J.BIORTECH.2022.127448
Kumar Awasthi, M., Paul, A., Kumar, V., Sar, T., Kumar, D., Sarsaiya, S., Liu, H., Zhang, Z.,
Binod, P., Sindhu, R., Kumar, V., & Taherzadeh, M. J. (2022). Recent trends and develop-
ments on integrated biochemical conversion process for valorization of dairy waste to value
added bioproducts: A review. Bioresource Technology, 344, 126193. https://doi.org/10.1016/J.
BIORTECH.2021.126193
Kumar, V., Sharma, N., Umesh, M., Selvaraj, M., Al-Shehri, B. M., Chakraborty, P., Duhan,
L., Sharma, S., Pasrija, R., Awasthi, M. K., Lakkaboyana, S. R., Andler, R., Bhatnagar,
A., & Maitra, S. S. (2022). Emerging challenges for the agro-industrial food waste utiliza-
tion: A review on food waste biorefinery. Bioresource Technology, 362, 127790. https://doi.
org/10.1016/J.BIORTECH.2022.127790
Ladakis, D., Stylianou, E., Ioannidou, S. M., Koutinas, A., & Pateraki, C. (2022). Biorefinery devel-
opment, techno-economic evaluation and environmental impact analysis for the conversion of
the organic fraction of municipal solid waste into succinic acid and value-added fractions.
Bioresource Technology, 354, 127172. https://doi.org/10.1016/J.BIORTECH.2022.127172
Le, T. S., Nguyen, P. D., Ngo, H. H., Bui, X. T., Dang, B. T., Diels, L., Bui, H. H., Nguyen, M. T.,
& Le Quang, D. T. (2022). Two-stage anaerobic membrane bioreactor for co-treatment of food
waste and kitchen wastewater for biogas production and nutrients recovery. Chemosphere, 309,
136537. https://doi.org/10.1016/J.CHEMOSPHERE.2022.136537
Lee, M. J., Song, J. H., & Hwang, S. J. (2009). Enhanced bio-energy recovery in a two-stage
hydrogen/methane fermentation process. Water Science and Technology, 59(11), 2137–2143.
https://doi.org/10.2166/WST.2009.236
Leong, Y. K., & Chang, J. S. (2022). Valorization of fruit wastes for circular bioeconomy: Current
advances, challenges, and opportunities. Bioresource Technology, 359, 127459. https://doi.
org/10.1016/J.BIORTECH.2022.127459
Li, D., Manu, M. K., Varjani, S., & Wong, J. W. C. (2022). Mitigation of NH3 and N2O emissions
during food waste digestate composting at C/N ratio 15 using zeolite amendment. Bioresource
Technology, 359, 127465. https://doi.org/10.1016/J.BIORTECH.2022.127465
Biomass Waste and Bioenergy Production: Challenges and Alternatives 65

Liang, J., Fang, W., Chang, J., Zhang, G., Ma, W., Nabi, M., Zubair, M., Zhang, R., Chen, L.,
Huang, J., & Zhang, P. (2022). Long-term rumen microorganism fermentation of corn stover
in vitro for volatile fatty acid production. Bioresource Technology, 358, 127447. https://doi.
org/10.1016/J.BIORTECH.2022.127447
Mahboubi, A., Uwineza, C., Doyen, W., De Wever, H., & Taherzadeh, M. J. (2020). Intensification
of lignocellulosic bioethanol production process using continuous double-staged immersed
membrane bioreactors. Bioresource Technology, 296, 122314. https://doi.org/10.1016/J.
BIORTECH.2019.122314
Mehariya, S., Goswami, R. K., Verma, P., Lavecchia, R., & Zuorro, A. (2021). Integrated approach
for wastewater treatment and biofuel production in microalgae biorefineries. Energies, 14(8),
2282. https://doi.org/10.3390/EN14082282
Mishra, A., Kumar, M., Bolan, N. S., Kapley, A., Kumar, R., & Singh, L. (2021). Multidimensional
approaches of biogas production and up-gradation: Opportunities and challenges. Bioresource
Technology, 338, 125514. https://doi.org/10.1016/J.BIORTECH.2021.125514
Monir, M. U., Aziz, A. A., Khatun, F., & Yousuf, A. (2020). Bioethanol production through syngas
fermentation in a tar free bioreactor using Clostridium butyricum. Renewable Energy, 157,
1116–1123. https://doi.org/10.1016/J.RENENE.2020.05.099
Narayanan, C. M., & Pandey, A. (2018). Studies on biodiesel synthesis using nanosilica immo-
bilised lipase in inverse fluidized bed bioreactors. Journal of Advances in Chemistry, 15(1),
6072–6086. https://doi.org/10.24297/jac.v15i1.7108
Paes, L. A. B., Bezerra, B. S., Deus, R. M., Jugend, D., & Battistelle, R. A. G. (2019). Organic
solid waste management in a circular economy perspective – A systematic review and
SWOT analysis. Journal of Cleaner Production, 239, 118086. https://doi.org/10.1016/J.
JCLEPRO.2019.118086
Pal, D. B., Tiwari, A. K., Mohammad, A., Prasad, N., Srivastava, N., Srivastava, K. R., Singh,
R., Yoon, T., Syed, A., Bahkali, A. H., & Gupta, V. K. (2022). Enhanced biogas produc-
tion potential analysis of rice straw: Biomass characterization, kinetics and anaerobic co-­
digestion investigations. Bioresource Technology, 358, 127391. https://doi.org/10.1016/J.
BIORTECH.2022.127391
Park, J. H., Chandrasekhar, K., Jeon, B. H., Jang, M., Liu, Y., & Kim, S. H. (2021). State-of-the-art
technologies for continuous high-rate biohydrogen production. Bioresource Technology, 320,
124304. https://doi.org/10.1016/J.BIORTECH.2020.124304
Piechota, G., Unpaprom, Y., Dong, C.-D., & Kumar, G. (2023). Recent advances in biowaste man-
agement towards sustainable environment. Bioresource Technology, 368, 128326. https://doi.
org/10.1016/J.BIORTECH.2022.128326
Rasapoor, M., Young, B., Brar, R., Sarmah, A., Zhuang, W. Q., & Baroutian, S. (2020). Recognizing
the challenges of anaerobic digestion: Critical steps toward improving biogas generation. Fuel,
261, 116497. https://doi.org/10.1016/J.FUEL.2019.116497
Raychaudhuri, A., Sahoo, R. N., & Behera, M. (2021). Application of clayware ceramic separa-
tor modified with silica in microbial fuel cell for bioelectricity generation during rice mill
wastewater treatment. Water Science and Technology, 84(1), 66–76. https://doi.org/10.2166/
WST.2021.213
Saini, A. K., Radu, T., Paritosh, K., Kumar, V., Pareek, N., Tripathi, D., & Vivekanand, V. (2021).
Bioengineered bioreactors: A review on enhancing biomethane and biohydrogen produc-
tion by CFD modeling. Bioengineered, 12(1), 6418–6433. https://doi.org/10.1080/2165597
9.2021.1972195
Saini, J. K., Himanshu, H., Kaur, A., & Mathur, A. (2022). Strategies to enhance enzymatic hydro-
lysis of lignocellulosic biomass for biorefinery applications: A review. Bioresource Technology,
360, 127517. https://doi.org/10.1016/J.BIORTECH.2022.127517
Saratale, R. G., Cho, S. K., Bharagava, R. N., Patel, A. K., Varjani, S., Mulla, S. I., Kim, D. S.,
Bhatia, S. K., Ferreira, L. F. R., Shin, H. S., & Saratale, G. D. (2022). A critical review on
biomass-based sustainable biorefineries using nanobiocatalysts: Opportunities, challenges,
66 A. Albahnasawi et al.

and future perspectives. Bioresource Technology, 363, 127926. https://doi.org/10.1016/J.


BIORTECH.2022.127926
Sim, Y. B., Jung, J. H., Baik, J. H., Park, J. H., Kumar, G., Rajesh Banu, J., & Kim,
S. H. (2021). Dynamic membrane bioreactor for high rate continuous biohydrogen produc-
tion from algal biomass. Bioresource Technology, 340, 125562. https://doi.org/10.1016/J.
BIORTECH.2021.125562
Sodhi, A. S., Sharma, N., Bhatia, S., Verma, A., Soni, S., & Batra, N. (2022). Insights on sustain-
able approaches for production and applications of value added products. Chemosphere, 286,
131623. https://doi.org/10.1016/J.CHEMOSPHERE.2021.131623
Sotiropoulos, A., Vourka, I., Erotokritou, A., Novakovic, J., Panaretou, V., Vakalis, S., Thanos, T.,
Moustakas, K., & Malamis, D. (2016). Combination of decentralized waste drying and SSF
techniques for household biowaste minimization and ethanol production. Waste Management,
52, 353–359. https://doi.org/10.1016/J.WASMAN.2016.03.047
Su, L., Zhou, F., Yu, M., Ge, R., He, J., Zhang, B., Zhang, Y., & Fan, J. (2020). Solid lipid
nanoparticles enhance the resistance of oat-derived peptides that inhibit dipeptidyl peptidase
IV in simulated gastrointestinal fluids. Journal of Functional Foods, 65, 103773. https://doi.
org/10.1016/J.JFF.2019.103773
Suriapparao, D. V., Gautam, R., & Rao Jeeru, L. (2022). Analysis of pyrolysis index and reac-
tion mechanism in microwave-assisted ex-situ catalytic co-pyrolysis of agro-­ residual
and plastic wastes. Bioresource Technology, 357, 127357. https://doi.org/10.1016/J.
BIORTECH.2022.127357
Tahir, N., Nadeem, F., Jabeen, F., Rani Singhania, R., Yaqub Qazi, U., Kumar Patel, A., Javaid,
R., & Zhang, Q. (2022). Enhancing biohydrogen production from lignocellulosic biomass
of Paulownia waste by charge facilitation in Zn doped SnO2 nanocatalysts. Bioresource
Technology, 355, 127299. https://doi.org/10.1016/J.BIORTECH.2022.127299
Teigiserova, D. A., Hamelin, L., & Thomsen, M. (2020). Towards transparent valorization of
food surplus, waste and loss: Clarifying definitions, food waste hierarchy, and role in the
circular economy. Science of the Total Environment, 706, 136033. https://doi.org/10.1016/J.
SCITOTENV.2019.136033
Thu Ha Tran, T., & Khanh Thinh Nguyen, P. (2022). Enhanced hydrogen production from water
hyacinth by a combination of ultrasonic-assisted alkaline pretreatment, dark fermentation, and
microbial electrolysis cell. Bioresource Technology, 357, 127340. https://doi.org/10.1016/J.
BIORTECH.2022.127340
Vakalis, S., Sotiropoulos, A., Moustakas, K., Malamis, D., Vekkos, K., & Baratieri, M. (2017).
Thermochemical valorization and characterization of household biowaste. Journal of
Environmental Management, 203, 648–654. https://doi.org/10.1016/J.JENVMAN.2016.04.017
Wainaina, S., Awasthi, M. K., Sarsaiya, S., Chen, H., Singh, E., Kumar, A., Ravindran, B., Awasthi,
S. K., Liu, T., Duan, Y., Kumar, S., Zhang, Z., & Taherzadeh, M. J. (2020). Resource recovery and
circular economy from organic solid waste using aerobic and anaerobic digestion technologies.
Bioresource Technology, 301, 122778. https://doi.org/10.1016/J.BIORTECH.2020.122778
Wang, H., Larson, R. A., Borchardt, M., & Spencer, S. (2019). Effect of mixing duration on biogas
production and methanogen distribution in an anaerobic digester. Environmental Technology,
42(1), 93–99. https://doi.org/10.1080/09593330.2019.1621951
Wang, X., Xu, T., de Andrade, M. J., Rampalli, I., Cao, D., Haque, M., Roy, S., Baughman, R. H.,
& Lu, H. (2021). The interfacial shear strength of carbon nanotube sheet modified carbon
fiber composites. In Conference proceedings of the society for experimental mechanics series
(pp. 25–32). https://doi.org/10.1007/978-­3-­030-­59542-­5_4/COVER
Xu, K., Lv, B., Huo, Y. X., & Li, C. (2018). Toward the lowest energy consumption and emis-
sion in biofuel production: Combination of ideal reactors and robust hosts. Current Opinion in
Biotechnology, 50, 19–24. https://doi.org/10.1016/J.COPBIO.2017.08.011
Xu, M., Yang, M., Sun, H., Gao, M., Wang, Q., & Wu, C. (2022). Bioconversion of biowaste into
renewable energy and resources: A sustainable strategy. Environmental Research, 214, 113929.
https://doi.org/10.1016/J.ENVRES.2022.113929
Biomass Waste and Bioenergy Production: Challenges and Alternatives 67

Zamri, M. F. M. A., Hasmady, S., Akhiar, A., Ideris, F., Shamsuddin, A. H., Mofijur, M., Fattah,
I. M. R., & Mahlia, T. M. I. (2021). A comprehensive review on anaerobic digestion of organic
fraction of municipal solid waste. Renewable and Sustainable Energy Reviews, 137, 110637.
https://doi.org/10.1016/J.RSER.2020.110637
Zhao, J., Li, Y., & Dong, R. (2021). Recent progress towards in-situ biogas upgrading tech-
nologies. Science of the Total Environment, 800, 149667. https://doi.org/10.1016/J.
SCITOTENV.2021.149667
Zhu, J., Jiao, N., Li, H., Xu, G., Zhang, H., & Xu, Y. (2022). p-Toluenesulfonic acid combined
with hydrogen peroxide-assisted pretreatment improves the production of fermentable sug-
ars from walnut (Juglans regia L.) shells. Bioresource Technology, 355, 127300. https://doi.
org/10.1016/J.BIORTECH.2022.127300
Enzyme-Mediated Strategies for Effective
Management and Valorization of Biomass
Waste

Usman Lawal Usman, Bharat Kumar Allam, and Sushmita Banerjee

1 Introduction

The world today faces a critical challenge in managing the mounting amounts of
biomass waste generated by various industries and human activities. Biomass waste,
which encompasses a varied range of organic substances derived from plants, ani-
mals, and microorganisms, poses significant environmental and economic concerns.
However, within this challenge lies an opportunity for innovative solutions that not
only address waste management but also unlock the untapped potential of biomass
as a valuable resource. In recent years, enzymes have emerged as powerful biocata-
lysts that hold great promise for the effective management and valorization of bio-
mass waste. These remarkable biomolecules, produced by living organisms, possess
exceptional catalytic capabilities and can facilitate a myriad of complex biochemi-
cal reactions (Mukherjee & Gupta, 2015). Leveraging the unique properties of
enzymes, researchers and scientists have been exploring novel strategies to convert
biomass waste into high-value products and biofuels, thereby mitigating environ-
mental pollution and creating a sustainable circular economy (Song et al., 2021).
This chapter focuses on the transformative role of enzyme-mediated strategies in
the management and valorization of biomass waste. It provides an in-depth
exploration of the recent state-of-the-art techniques and highlights the potential

U. L. Usman
Department of Biology, Umaru Musa Yar’adua University, Katsina, Nigeria
Department of Environmental Sciences, Sharda University, Greater Noida, India
B. K. Allam
Department of Chemistry, Faculty of Basic Sciences, Rajiv Gandhi University
(A Central University), Doimukh, Arunachal Pradesh, India
S. Banerjee (*)
Department of Environmental Sciences, Sharda University, Greater Noida, India
e-mail: sushmita.banerjee@sharda.ac.in

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 69


A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_4
70 U. L. Usman et al.

applications across various industries. By delving into the fundamental principles of


enzymatic reactions and their interactions with biomass substrates, this chapter
aims to elucidate the underlying mechanisms that enable the efficient conversion of
waste into valuable resources.

1.1 Background on Biomass Waste and Its


Environmental Impact

Biomass waste is generated from a wide range of sources, including agriculture,


forestry, food processing, and municipal activities. It encompasses various organic
materials such as crop residues, wood chips, sawdust, animal manure, food scraps,
and wastewater sludge. The accumulation of biomass waste poses substantial envi-
ronmental challenges because of its decomposition process, which releases green-
house gases such as carbon dioxide and methane into the atmosphere (Amalina
et al., 2022). Additionally, the improper handling and disposal of biomass waste can
result in soil and water pollution, affecting human health and ecosystems. The envi-
ronmental impact of biomass waste extends beyond its contribution to greenhouse
gas emissions. When biomass waste is incinerated or decomposes anaerobically in
landfills, it releases pollutants for instance sulfur dioxide, nitrogen oxides, and vola-
tile organic compounds. These pollutants contribute to air pollution and can have
detrimental effects on air quality and human respiratory health. Furthermore, the
leaching of organic and inorganic compounds from biomass waste into soil and
water bodies can contaminate groundwater and surface water, disrupting ecosys-
tems and endangering aquatic life (Abdel-Shafy & Mansour, 2018). The sheer vol-
ume of biomass waste generated globally emphasizes the urgency to address its
environmental impact. In 2018, the Food and Agriculture Organization (FAO) stated
that approximately 2.01 billion metric tons of biomass waste were generated world-
wide (Ogbu & Okechukwu, 2023). This number is projected to increase due to
population growth, urbanization, and changes in consumption patterns. Recognizing
the detrimental effects of unmanaged biomass waste, governments, industries, and
research communities are actively seeking effective strategies to mitigate its envi-
ronmental impact. Emphasizing the principles of the circular economy, these efforts
aim to minimize waste generation, maximize resource recovery, and reduce reliance
on nonrenewable resources (Sasmoko et al., 2022). Enzyme-mediated strategies
have emerged as promising solutions for the effective management and valorization
of biomass waste, enabling its conversion into valuable products and reducing its
environmental footprint. By harnessing the power of enzymes, biomass waste can
be transformed into biofuels, biochemicals, biopolymers, and other valuable inter-
mediates (Nargotra et al., 2023). Enzymes break down complex organic compounds
present in biomass waste into simpler forms, facilitating their conversion into usable
and marketable products. This approach offers numerous advantages as compared
to the traditional methods, which include a reduction in energy consumption, milder
reaction conditions, and improved selectivity.
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 71

1.2 Importance of Effective Management


and Valorization Strategies

Effective management and valorization strategies for biomass waste are of para-
mount importance due to several key reasons:
(i) Environmental Sustainability: Biomass waste, if not managed properly, can
have significant environmental impacts, which include air pollution such as
greenhouse gas emissions and contamination of water and soil. By implement-
ing effective management strategies, such as proper collection, sorting, and
treatment, biomass waste can be minimized, and its negative environmental
consequences can be mitigated. Valorization strategies, on the other hand,
enable biomass waste conversion into useful products, reducing the need for
resource extraction and minimizing the overall environmental footprint (Abdel-­
Shafy & Mansour, 2018).
(ii) Resource Conservation: Biomass waste contains valuable organic compounds
that can be utilized as a renewable resource. Effective management and valori-
zation strategies enable the recovery and utilization of these resources, decreas-
ing the dependency on nonrenewable resources through the conversion of
biomass waste into biofuels, biochemicals, and other useful products, and we
can conserve finite resources, decrease reliance on fossil fuels, and promote a
more sustainable and circular economy (Okafor et al., 2022).
(iii) Waste Reduction and Recycling: Biomass waste, if left unmanaged, can accu-
mulate in landfills, occupying valuable land and emitting greenhouse gases
during decomposition. By implementing efficient management strategies, such
as waste reduction, recycling, and composting, the volume of biomass waste
sent to landfills can be minimized. Additionally, valorization strategies allow
for the transformation of biomass waste into usable products, reducing the
need for virgin materials and contributing to waste diversion from landfills
(Zhou & Wang, 2020).
(iv) Economic Opportunities: Effective management and valorization of biomass
waste can create new economic opportunities. The development of biomass
waste processing facilities, biorefineries, and related industries can generate
jobs and stimulate economic growth. Valorization strategies can lead to the
development of bio-based products, such as biofuels, bioplastics, and bio-
chemicals, which have a growing market demand. By tapping into the potential
of biomass waste, countries and industries can diversify their economies,
enhance competitiveness, and foster sustainable development (Nizami
et al., 2017).
(v) Climate Change Mitigation: The management and valorization of biomass
waste play a crucial role in mitigating climate change. Biomass waste, if left
untreated, emits greenhouse gases during decomposition or incineration. By
implementing effective valorization strategies, such as anaerobic digestion or
thermochemical conversion, biomass waste can be transformed into biofuels
72 U. L. Usman et al.

and other bio-based products. These biofuels can replace fossil fuels, thereby
lowering the emissions of greenhouse gases and contributing to a low-carbon
economy transition (Awogbemi & Von Kallon, 2022).

1.3 Role of Enzymes in Biomass Waste Processing

Enzymes play a crucial role in biomass waste processing, offering efficient and eco-­
friendly solutions for the complex organic compounds’ conversion into useful prod-
ucts. Enzymes are biocatalysts, typically proteins, that increase biochemical
reactions by reducing the activation energy needed for the occurring reaction
(Robinson, 2015). They possess remarkable specificity, acting on specific substrates
and catalyzing specific reactions, making them highly suitable for biomass waste
processing as depicted in Fig. 1.
(i) Enzymatic Degradation: Biomass waste consists of various complex poly-
mers, such as lignin, hemicellulose, and cellulose. Enzymes involved in the
degradation of biomass including cellulases, ligninases, and hemicellulases
break down these complex polymers into simpler units. For example, cellu-
lases hydrolyze cellulose into glucose, hemicellulose is broken down by
hemicellulases into various sugars, and ligninases convert lignin into simpler
aromatic compounds (Kumar & Chandra, 2020). Enzymatic degradation of

Fig. 1 Overview role of enzymes in biomass waste processing


Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 73

biomass waste enables the release of valuable sugars and other intermediates
that can be further utilized in various downstream processes.
(ii) Enzyme-Assisted Pretreatment: Biomass waste often requires pretreatment to
enhance its enzymatic digestibility. Enzyme-assisted pretreatment encom-
passes the use of specific enzymes to modify the biomass structure and com-
position, making it more susceptible to enzymatic degradation. For instance,
enzymes such as ligninases can selectively remove or modify lignin, increas-
ing the hemicellulose and cellulose accessibility to subsequent enzymatic
hydrolysis (Sharma et al., 2023a). Enzyme-assisted pretreatment methods
offer advantages over traditional chemical pretreatment methods by operating
under milder conditions and reducing the formation of toxic by-products.
(iii) Enzymatic Hydrolysis: Enzymatic hydrolysis is a key step in biomass waste
processing, where enzymes break down complex carbohydrates into simple
sugars. Cellulases and hemicellulases act synergistically to cleave the bonds
in the glycosidic present in the cellulose and hemicellulose respectively. The
resulting sugars, such as xylose and glucose, can be used in biofuels, bio-
chemicals, and other useful materials production. Enzymatic hydrolysis
offers several advantages, including higher selectivity, lower energy require-
ments, and milder reaction conditions as compared to traditional acid or
enzymatic processes (Houfani et al., 2020).
(iv) Enzyme-Mediated Conversion: Enzymes also play a critical role in biomass
waste conversion into value-added products. Various enzymes are involved in
specific conversion pathways, such as biofuel production (e.g., cellulases for
ethanol production) or biochemicals (e.g., enzymes for the production of plat-
form chemicals such as succinic acid or lactic acid). Enzymes enable the
transformation of biomass waste into a variety of valuable products, compris-
ing biofuels, biopolymers, organic acids, and enzymes themselves (Fülöp &
Ecker, 2020).
(v) Enzyme Immobilization: Enzyme immobilization techniques are employed to
increase enzymatic stability and reusability in biomass waste processing.
Immobilized enzymes are attached or confined to a support material, which
can be solid or porous. Immobilization improves enzyme performance, allow-
ing for their repeated use over multiple reaction cycles. Immobilized enzymes
are particularly beneficial for large-scale biomass waste processing, offering
advantages such as improved operational stability, simplified enzyme separa-
tion from the reaction mixture, and reduced enzyme cost (Homaei et al.,
2013). Enzyme-mediated strategies provide several advantages in biomass
waste processing, including higher reaction specificity, milder reaction condi-
tions, reduced consumption of energy, and generation of minimal toxic by-­
products. The ongoing advancements in enzyme engineering, enzyme
immobilization techniques, and process optimization are further enhancing
the efficiency and scalability of enzymatic processes. Enzymes have the
potential to revolutionize biomass waste processing by enabling the utiliza-
tion of this abundant and renewable resource for sustainable and value-added
product production (Yaashikaa et al., 2022).
74 U. L. Usman et al.

(vi) Enzyme Engineering and Optimization: Enzyme engineering techniques,


such as protein engineering, allow for the modification and optimization of
enzymes to improve their catalytic activity, substrate specificity, stability, and
resistance to inhibitors. Through protein engineering, enzymes can be tai-
lored to better suit specific biomass waste compositions or to perform under
specific process conditions (Singh et al., 2013). Optimization techniques,
such as enzyme loading optimization and reaction parameter optimization,
help maximize the efficiency and yield of enzymatic processes. These
advancements in enzyme engineering and optimization contribute to the syn-
thesis of more efficient and cost-effective enzyme-mediated biomass waste
processing methods.
(vii) Synergistic Enzyme Systems: Biomass waste is a complex mixture of poly-
mers that require the concerted action of multiple enzymes for efficient con-
version. Synergistic enzyme systems, composed of different enzymes working
together, are employed to tackle the complexity of biomass waste (Zamora
Zamora et al., 2020). These enzyme systems involve the simultaneous or
sequential action of complementary enzymes, such as cellulases, hemicellu-
lases, and ligninases, to effectively degrade and convert the various compo-
nents of biomass waste. By harnessing the synergistic interactions between
enzymes, the overall efficiency of biomass waste processing can be signifi-
cantly enhanced.
(viii) Enzyme Technology Integration: Enzyme-mediated strategies can be com-
bined with other techniques to further enhance the efficiency and sustainabil-
ity of biomass waste processing. For example, enzymatic processes can be
combined with microbial fermentation to convert sugars derived from bio-
mass waste into biochemicals or biofuels. Additionally, enzymatic processes
can be incorporated with chemical catalysis or thermochemical conversion
techniques to enable a wider range of biomass waste valorization pathways.
The integration of enzyme technology with other processes offers opportuni-
ties for process intensification, improved product selectivity, and higher over-
all process efficiency (Singh et al., 2022).

2 Enzymatic Degradation of Biomass Waste

Biomass waste comprises a wide range of organic materials, including lignocellu-


losic biomass from agricultural residues (e.g., wheat straw, corn stover), forestry
by-products (e.g., sawdust, wood chips), and food waste (Jekayinfa et al., 2020).
These biomass sources are rich in complex polymers such as hemicellulose, cellu-
lose, and lignin, which can be targeted for enzymatic degradation. Enzymatic deg-
radation is particularly effective for lignocellulosic biomass, as it can unlock the
valuable components trapped within the complex structure of these materials.
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 75

2.1 Types of Biomass Waste Suitable


for Enzymatic Degradation

Enzymatic degradation is a powerful approach for converting complex organic


compounds in biomass waste into simpler, more valuable products. Various types of
biomass waste are suitable for enzymatic degradation, primarily focusing on ligno-
cellulosic biomass (Zhou & Wang, 2020). Table 1 provides some examples.

2.2 Major Enzymes Involved in Biomass Degradation

Enzymatic degradation of biomass waste relies on the activity of various enzymes


that can specifically target the complex polymers present in biomass. These enzymes
play a crucial role in breaking down cellulose, hemicellulose, and lignin, which are
the major components of biomass waste (Houfani et al., 2020). Figure 2 shows
some basic enzymes used in biomass degradation.
(i) Cellulases
Cellulases are various groups of enzymes that catalyze the hydrolysis of
cellulose, a linear polymer of glucose units. They consist of three main catego-
ries of enzymes: exoglucanases (also known as cellobiohydrolases), endoglu-
canases, and β-glucosidases. The endoglucanases randomly cleave internal
bonds within the cellulose chain, generating shorter cellulose fragments.
Exoglucanases act on the ends of cellulose chains, releasing cellobiose units.
β-glucosidases further hydrolyze cellobiose into glucose. Cellulases are cru-
cial for breaking down cellulose into fermentable sugars that can be utilized for

Table 1 Types of biomass waste suitable for enzymatic degradation


Major biomass
Biomass waste type components Enzymes involved
Agricultural residues such as crop stalks, Cellulose, Cellulases,
husks, and straw hemicellulose, lignin hemicellulases,
ligninases
Forest residues including branches, leaves, and Cellulose, Cellulases,
bark hemicellulose, lignin hemicellulases,
ligninases
Energy crops such as switchgrass or Cellulose, Cellulases,
miscanthus hemicellulose, lignin hemicellulases,
ligninases
Food processing waste including vegetable Starch, cellulose, Amylases, cellulases,
peels, fruit pulp, and by-products from food proteins proteases
manufacturing
Municipal solid waste comprises a mixture of Organic matter, Cellulases, amylases,
organic matter and cellulosic materials cellulosic materials proteases
76 U. L. Usman et al.

Fig. 2 Overview of some major enzymes involved in biomass degradation

the production of biofuel or other biochemical applications (Jayasekara &


Ratnayake, 2019).
(ii) Hemicellulases
Hemicellulases are varied groups of enzymes used in the degradation of
hemicellulose, a branched polymer consisting of various sugar units.
Hemicellulases include enzymes such as xylanases, mannanases, xylogluca-
nases, and arabinases, which target different types of hemicellulose found in
biomass waste. These enzymes hydrolyze the glycosidic bonds within hemi-
cellulose, releasing a range of sugar monomers, for example, mannose, xylose,
galactose, and arabinose. Hemicellulases are important for the complete utili-
zation of biomass waste and the production of diverse value-added products
(Jayasekara & Ratnayake, 2019).
(iii) Ligninases
Ligninases are enzymes that are involved in the breaking down of lignin, the
complex and highly recalcitrant polymer that provides rigidity and protection
to plant cell walls. Ligninases encompass different enzyme groups, comprising
laccases, manganese peroxidases, and lignin peroxidases. These enzymes
work through oxidative mechanisms, breaking down the aromatic structure of
lignin. Ligninases play a vital role in lignin degradation, facilitating the access
of other enzymes to cellulose and hemicellulose components within biomass
waste (Kumar & Chandra, 2020).
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 77

(iv) Other Enzymes


Besides cellulases, hemicellulases, and ligninases, other enzymes may also
contribute to biomass degradation. For instance, pectinases are enzymes that
degrade pectin, a complex polysaccharide found in the plants’ primary cell
walls. Pectinases break down pectin into simpler sugar units, such as galact-
uronic acid. Enzymes such as proteases and amylases may also be used in the
degradation of starch and proteins present in certain biomass waste streams.
The selection of enzymes for biomass degradation is based on the specific
composition and structure of the biomass waste being processed. Commercial
enzyme cocktails, which consist of a mixture of different enzymes, are avail-
able and often used to ensure the efficient degradation of complex biomass
substrates (Jayasekara & Ratnayake, 2019). Moreover, ongoing advancements
in enzyme engineering and protein modification techniques allow for the opti-
mization of enzyme properties, such as improved activity, stability, and sub-
strate specificity, to enhance their performance in biomass degradation
processes.

2.3 Mechanisms of Enzymatic Degradation and Key Factors


Affecting Enzyme Activity

Enzymatic degradation of biomass waste involves specific mechanisms by which


enzymes break down complex polymers into simpler units. The efficiency of enzy-
matic degradation is influenced by various factors that affect enzyme activity.
Understanding these mechanisms and key factors is essential for optimizing enzy-
matic processes for biomass waste degradation (Chen et al., 2020). Table 2 high-
lights some enzymes involved in biomass degradation, and the mechanisms of
enzymatic degradation and the key factors that impact enzyme activity are dis-
cussed below.

2.3.1 Mechanisms of Enzymatic Degradation

(a) Hydrolytic Mechanism: Many enzymes involved in biomass degradation, such


as cellulases and hemicellulases, operate through a hydrolytic mechanism.
They break down the glycosidic bonds within the polymer structure by adding
water molecules, resulting in the bond cleavage and the release of smaller sugar
units (Estela & Luis, 2013). This hydrolysis process enables the conversion of
complex carbohydrates into fermentable sugars.
(b) Oxidative Mechanism: Ligninases, including manganese peroxidases, lignin
peroxidases, and laccases, employ an oxidative mechanism to degrade lignin.
These enzymes utilize reactive oxygen species (ROS) to break down the aro-
matic structure of lignin, resulting in the production of smaller lignin fragments.
Oxidative degradation of lignin involves the formation of free radicals and the
subsequent cleavage of chemical bonds (Estela & Luis, 2013).
78 U. L. Usman et al.

Table 2 Enzymes involved in biomass degradation


Biomass
component
Enzyme type targeted Function
Cellulases are enzymes that target cellulose, a major Cellulose Hydrolyze
component of biomass. Their function is to hydrolyze the cellulose
glycosidic bonds within cellulose, breaking it down into
smaller sugar units, such as glucose
Hemicellulases are enzymes that act on hemicellulose, Hemicellulose Hydrolyze
another important component of biomass. They hydrolyze hemicellulose
the complex structure of hemicellulose, releasing various
sugar units, such as xylose, mannose, and arabinose
Ligninases are enzymes employed in the degradation of Lignin Degrade lignin
lignin, a complex polymer found in the cell walls of plants.
Their function is to degrade and break down the intricate
structure of lignin, enabling the separation of lignin from
other biomass components
Amylases are enzymes that target starch, a carbohydrate Starch Hydrolyze
reserve in many plant-based materials. They catalyze the starch
hydrolysis of starch into simpler sugar units, such as
maltose and glucose
Proteases also known as proteolytic enzymes act on proteins Proteins Hydrolyze
present in biomass. Their function is to hydrolyze peptide proteins
bonds, breaking down proteins into amino acids or smaller
peptides

2.3.2 Key Factors Affecting Enzyme Activity

(a) pH: Enzyme activity is influenced by pH, as enzymes have an optimal pH range
at which they exhibit the highest activity. Different enzymes involved in bio-
mass degradation have varying pH optima. For example, cellulases typically
have an optimal pH range between 4.5 and 6.5, while ligninases may have opti-
mal pH values ranging from acidic to alkaline conditions. Maintaining the
appropriate pH for the specific enzymes used is crucial for achieving optimal
enzymatic activity (Robinson, 2015).
(b) Temperature: Enzyme activity is also temperature-dependent, with each enzyme
having an optimal temperature range. Higher temperatures generally enhance
enzymatic activity, but excessive heat can lead to enzyme denaturation and loss
of activity. Enzymes involved in biomass degradation often exhibit optimal
activity within a moderate temperature range, typically between 40 °C and
70 °C, depending on the specific enzyme. Careful control of temperature is
necessary to maintain enzyme stability and activity during biomass degradation
(Robinson, 2015).
(c) Substrate Concentration: The concentration of the substrate, such as cellulose
or hemicellulose, affects enzyme activity. Generally, enzyme activity increases
with increasing substrate concentration until a saturation point is reached.
Beyond this point, further increases in substrate concentration may not signifi-
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 79

cantly enhance enzyme activity. Optimizing the substrate concentration is


important to ensure efficient enzymatic degradation without excessive substrate
wastage (German et al., 2011).
(d) Enzyme-Substrate Affinity: Enzyme-substrate affinity denotes the strength of
the interaction between the enzyme and its substrate. The affinity determines
the efficiency with which the enzyme can bind to and catalyze the degradation
of the substrate. Enzymes with high affinity exhibit stronger binding to the
substrate, leading to increased degradation efficiency. Modifying enzyme-­
substrate affinity through enzyme engineering techniques can enhance the cata-
lytic efficiency of enzymes involved in biomass degradation (Robinson, 2015).
(e) Enzyme Stability: Enzyme stability refers to the ability of the enzyme to main-
tain its activity over time under specific conditions. Stability is influenced by
factors such as temperature, pH, and the presence of inhibitors. Enhanced sta-
bility allows enzymes to retain their activity during prolonged degradation
­processes, minimizing the need for frequent enzyme replacement and improv-
ing overall process efficiency (German et al., 2011).
(f) Enzyme Loading: Enzyme loading refers to the amount of enzyme used in the
biomass degradation process. The appropriate enzyme loading is crucial for
achieving efficient degradation without excessive enzyme costs. Optimization
of enzyme loading involves balancing the degradation efficiency with the cost-­
effectiveness of the process (Guo et al., 2023).

3 Enzyme Engineering for Improved Degradation Efficiency

Enzyme engineering plays a crucial role in improving the degradation efficiency of


enzymes involved in biomass waste processing. Recent advancements in enzyme
engineering techniques have enabled the modification and optimization of enzymes
to enhance their catalytic activity, substrate specificity, stability, and resistance to
inhibitors (Zhu et al., 2022). These advancements have significantly contributed to
improving the efficiency of enzymatic degradation processes. Figure 3 shows some
recent advancements in enzyme engineering for improved degradation efficiency.

3.1 Challenges and Limitations of Enzymatic Degradation

While enzymatic degradation holds great promise for the efficient management and
valorization of biomass waste, several challenges and limitations need to be
addressed. Understanding these challenges is crucial for developing effective strate-
gies and technologies to overcome them. Here are some key challenges and limita-
tions associated with enzymatic degradation:
80 U. L. Usman et al.

Fig. 3 An overview of some recent developments in enzyme engineering processes

(i) Recalcitrant Nature of Biomass Waste: Biomass waste, particularly lignocel-


lulosic materials, possesses a complex and recalcitrant structure. The presence
of lignin, a highly resistant polymer, hinders the access of enzymes to cellu-
lose and hemicellulose, making degradation more challenging. Lignin also
acts as a physical barrier, limiting the efficiency of enzymatic degradation.
Strategies such as pretreatment methods (chemical, physical, or biological)
are employed to overcome this challenge by disrupting the lignin structure and
improving enzyme accessibility to the target polymers (Mishra et al., 2018).
(ii) Substrate Heterogeneity and Variability: Biomass waste from different sources
can vary significantly in terms of composition, structure, and properties. This
heterogeneity poses a challenge for enzymatic degradation, as enzymes need
to be tailored to suit specific biomass compositions. Different biomass sources
may require different enzyme cocktails or specific enzymes with enhanced
substrate specificity. Developing enzyme systems that can efficiently degrade
a wide range of biomass feedstocks remains a challenge (Sweeney &
Xu, 2012).
(iii) Enzyme Production Costs: Enzyme production costs can be a significant bar-
rier to large-scale enzymatic degradation processes. Many enzymes used in
biomass degradation, especially those with high specificities, are costly to pro-
duce. The scale-up of enzyme production and purification processes needs to
be optimized to ensure cost-effectiveness. Strategies such as enzyme immobi-
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 81

lization and enzyme recycling can help reduce enzyme consumption and mini-
mize overall production costs (Sakhuja et al., 2021).
(iv) Enzyme Stability and Inhibition: Enzymes used in biomass degradation may
face stability challenges due to harsh process conditions, such as high tem-
peratures and extreme pH. These conditions can cause enzyme denaturation or
reduced activity over time. Moreover, the presence of inhibitors, such as
lignin-­derived compounds or other by-products from biomass degradation,
can inhibit enzyme activity. Enhancing enzyme stability and developing
enzymes with improved tolerance to inhibitors are areas of ongoing research
(Robinson, 2015).
(v) Process Integration and Scale-Up: Integrating enzymatic degradation into
existing biomass waste processing infrastructures and scaling up enzymatic
processes pose significant challenges. The compatibility of enzymatic pro-
cesses with other process steps, such as pretreatment, fermentation, or down-
stream separation, needs to be addressed for efficient integration. Furthermore,
the development of cost-effective and scalable reactor systems and process
optimization strategies is essential to enable large-scale enzymatic degrada-
tion of biomass waste (Balan, 2014).
(vi) Techno-Economic Considerations: Evaluating the techno-economic feasibility
of enzymatic degradation processes is crucial for their practical implementa-
tion. The cost-effectiveness of enzymatic processes should be carefully
assessed, considering factors such as enzyme production costs, substrate avail-
ability and costs, downstream processing requirements, and market demand
for the resulting products. The economics of biomass waste degradation need
to be competitive with alternative waste management and valorization
approaches (Robinson, 2015).
(vii) Regulatory and Public Acceptance: The adoption of enzymatic degradation
processes for biomass waste management may face regulatory challenges and
public acceptance issues. Ensuring the safety and environmental impact of the
enzymatic processes, addressing concerns about genetically modified organ-
isms (GMOs) used for enzyme production, and complying with regulations
governing waste management and product quality are important consider-
ations (Wesseler et al., 2023).

4 Enzyme-Assisted Pretreatment of Biomass Waste

Enzyme-assisted pretreatment is a critical step in the effective management and


valorization of biomass waste. It involves the application of enzymes to modify the
structure and composition of biomass waste, enhancing its susceptibility to subse-
quent enzymatic degradation. Enzyme-assisted pretreatment strategies can improve
the accessibility of enzymes to target polymers and increase the overall efficiency of
82 U. L. Usman et al.

Table 3 Enzymatic pretreatment methods and advantages


Pretreatment method Advantages
Acid pretreatment involves the biomass treatment with acids such as Effective
hydrochloric acid or sulfuric acid. It offers effective delignification, delignification, high
breaking down the lignin component of biomass, which improves the sugar release
accessibility of cellulose and hemicellulose to subsequent enzymatic
hydrolysis. Acid pretreatment also leads to high sugar release due to the
breakdown of hemicellulose into fermentable sugars
Alkaline pretreatment employs alkaline solutions such as sodium Efficient
hydroxide or ammonium hydroxide to treat biomass. It efficiently hemicellulose
removes hemicellulose from the biomass matrix, enhancing the removal, reduced
accessibility of cellulose for enzymatic hydrolysis. Alkaline pretreatment inhibitor formation
also reduces the formation of inhibitors that can hinder enzymatic
reactions, improving overall process efficiency
Steam explosion is a pretreatment method that subjects biomass to Disrupts biomass
high-pressure steam followed by rapid decompression. This process structure enhances
disrupts the biomass structure, resulting in the swelling of cellulose and enzyme accessibility
the creation of porous structures. It enhances the accessibility of
enzymes to cellulose, improving enzymatic hydrolysis efficiency
Liquid hot water pretreatment involves treating biomass with water Mild conditions, low
under high temperature and pressure. It solubilizes and removes energy requirement
hemicellulose from the biomass while maintaining the integrity of
cellulose. Liquid hot water pretreatment operates under milder
conditions compared to other methods, reducing energy requirements
and facilitating subsequent enzymatic hydrolysis
Organosolv pretreatment utilizes organic solvents, such as ethanol or Effective lignin
methanol, in combination with dilute acid or base to treat biomass. This removal, high sugar
method effectively removes lignin from biomass while preserving yield
cellulose and hemicellulose components. Organosolv pretreatment
enables high sugar yields and produces high-quality lignin for various
applications

biomass degradation (Sharma et al., 2023b). Table 3 explores the key aspects and
approaches of enzyme-assisted pretreatment in biomass waste processing.

4.1 Synergistic Effects Between Pretreatment


and Enzymatic Degradation

Enzyme-assisted pretreatment can synergistically enhance the efficiency of subse-


quent enzymatic degradation. Pretreatment methods modify the structure and com-
position of biomass waste, increasing the exposure of cellulose and hemicellulose
to enzymes. This leads to improved enzymatic hydrolysis rates, reduced enzyme
requirements, and enhanced overall degradation efficiency (Zhao et al., 2021). The
combination of pretreatment and enzymatic degradation offers a more effective
approach for biomass waste valorization compared to enzymatic degradation alone.
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 83

4.2 Optimization of Pretreatment Conditions

The optimization of pretreatment conditions is crucial to achieving the desired out-


comes in enzyme-assisted pretreatment. Parameters such as pretreatment tempera-
ture, duration, pH, enzyme loading, and substrate concentration need to be carefully
controlled and optimized to maximize the efficiency of pretreatment (Sharma et al.,
2023a). Optimization aims to achieve the highest possible delignification and hemi-
cellulose removal while minimizing the degradation of cellulose and reducing the
formation of inhibitory compounds. Understanding biomass composition and char-
acteristics is essential for developing pretreatment strategies tailored to specific bio-
mass waste sources.

4.3 Enzymatic Pretreatment Methods and Their Advantages

Enzymatic pretreatment methods utilize specific enzymes to modify the structure


and composition of biomass waste, enhancing its susceptibility to subsequent con-
version processes. These methods offer several advantages over other pretreatment
approaches, such as physical or chemical methods. Here, we explore enzymatic
pretreatment methods and their associated advantages:

4.3.1 Cellulase-Based Enzymatic Pretreatment

Cellulase-based enzymatic pretreatment involves the application of cellulases, a


group of enzymes that specifically target cellulose, to modify the cellulose structure
in biomass waste. Cellulases can hydrolyze the glycosidic bonds within cellulose,
leading to the disruption of cellulose fibers and the formation of smaller cellulose
fragments. This enzymatic pretreatment method offers the following advantages
(Jayasekara & Ratnayake, 2019):
(a) Specificity: Cellulases exhibit high specificity for cellulose, selectively target-
ing and degrading cellulose polymers. This allows for the precise modification
of cellulose without significantly affecting other components, such as hemicel-
lulose or lignin.
(b) Mild Operating Conditions: Cellulase-based enzymatic pretreatment operates
under mild conditions, typically at moderate temperatures and near-neutral
pH. This reduces energy consumption and minimizes the generation of harmful
by-products, making the process environmentally friendly.
(c) Enzyme Recycling: Enzymes used in cellulase-based enzymatic pretreatment
can be recovered and recycled for repeated use. This helps to minimize enzyme
costs and improves the overall economic feasibility of the process.
84 U. L. Usman et al.

(d) Compatibility with Downstream Processes: Enzymatic pretreatment using cel-


lulases can be seamlessly integrated with subsequent enzymatic hydrolysis or
fermentation processes. The compatibility between the enzymes used in pre-
treatment and downstream processes ensures the efficient conversion of cellu-
lose into desired products, such as sugars, biofuels, or biochemicals.

4.3.2 Ligninase-Based Enzymatic Pretreatment

Ligninase-based enzymatic pretreatment involves the application of ligninolytic


enzymes, for example, manganese peroxidases, lignin peroxidases, or laccases to
modify the lignin structure in biomass waste. These enzymes can break down lig-
nin, facilitating the subsequent conversion of cellulose and hemicellulose compo-
nents. Enzymatic pretreatment using ligninases offers the following advantages
(Kumar & Chandra, 2020):
(a) Selective Lignin Degradation: Ligninases specifically target lignin, leaving cel-
lulose and hemicellulose relatively intact. This selectivity allows for the con-
trolled modification of lignin, enhancing the accessibility of cellulose and
hemicellulose for subsequent enzymatic hydrolysis or fermentation.
(b) Reduced Formation of Inhibitory Compounds: Ligninase-based enzymatic pre-
treatment leads to the breakdown of lignin, reducing the formation of inhibitory
compounds that can negatively impact downstream processes. By reducing the
inhibitory effects of lignin-derived compounds, enzymatic pretreatment
enhances the efficiency and yields of subsequent conversion processes.
(c) Enzyme Compatibility and Synergy: Ligninases can work synergistically with
other enzymes, such as cellulases or hemicellulases, in a cocktail approach. The
combined action of ligninases and other enzymes can enhance the overall effi-
ciency of biomass degradation, resulting in higher sugar yields and improved
product quality.
(d) Environmental Friendliness: Enzymatic pretreatment using ligninases operates
under mild conditions, requiring moderate temperatures and near-neutral
pH. This minimizes the use of harsh chemicals and reduces the environmental
impact associated with conventional chemical pretreatment methods.

4.3.3 Effect of Enzyme-Assisted Pretreatment on Biomass Structure


and Composition

Enzyme-assisted pretreatment plays a significant role in modifying the structure and


composition of biomass waste, enhancing its susceptibility to subsequent conver-
sion processes. The application of specific enzymes during pretreatment brings
about several changes in biomass, leading to improved accessibility, increased
enzymatic hydrolysis efficiency, and higher yields of desired products (Ramos
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 85

et al., 2020). Below are some of the effects of enzyme-assisted pretreatment on


biomass structure and composition:
(i) Disruption of Biomass Structure
Enzyme-assisted pretreatment leads to the disruption of the complex struc-
ture of biomass waste, particularly lignocellulosic materials. The enzymes
employed in pretreatment act on specific components of biomass, such as cel-
lulose, hemicellulose, or lignin, resulting in structural modifications (Abraham
et al., 2020). The key effects of enzyme-assisted pretreatment on biomass
structure include the following:
(a) Cellulose Modification: Enzymatic pretreatment, especially with cellu-
lases, breaks down cellulose chains into smaller fragments, leading to the
disruption of the crystalline structure of cellulose. This increases the
accessible surface area and accessibility of cellulose for subsequent enzy-
matic hydrolysis.
(b) Hemicellulose Degradation: Enzymes targeting hemicellulose, such as
hemicellulases, hydrolyze the glycosidic bonds within hemicellulose,
resulting in the breakdown of the hemicellulosic polysaccharides. This
modification of hemicellulose structure improves the exposure of cellulose
and other components, allowing for more efficient enzymatic
degradation.
(c) Lignin Modification: Enzymes, such as ligninases or peroxidases, modify
the lignin component of biomass waste. They catalyze oxidative reactions
or cleave specific bonds within lignin, leading to the partial breakdown or
modification of lignin polymers. This alters the lignin structure, loosening
its association with cellulose and hemicellulose and improving its acces-
sibility for subsequent conversion processes.
(ii) Increased Porosity and Surface Area
Enzyme-assisted pretreatment increases the porosity and surface area of
biomass waste, facilitating the access of enzymes or microorganisms to the
target polymers. The effects of enzyme-assisted pretreatment on porosity and
surface area include the following:
(a) Increased Porosity: The breakdown of cellulose, hemicellulose, and lignin
during pretreatment creates pores and voids within the biomass structure.
This increases the overall porosity of biomass, enhancing the diffusion of
enzymes or microorganisms and facilitating the accessibility of enzymes
to the target polymers (Sharma et al., 2023a).
(b) Increased Surface Area: Enzymatic pretreatment breaks down biomass
components into smaller fragments, exposing a larger surface area for
enzymatic hydrolysis or microbial fermentation. The increased surface
area improves the contact between enzymes or microorganisms and the
biomass substrate, leading to enhanced conversion rates and higher yields
of desired products.
86 U. L. Usman et al.

(iii) Removal of Inhibitory Compounds


Enzyme-assisted pretreatment can remove or modify inhibitory compounds
present in biomass waste, improving the efficiency of subsequent conversion
processes. The effects of enzyme-assisted pretreatment on inhibitory com-
pounds include the following:
(a) Removal of Lignin-Derived Inhibitors: Lignin degradation during pretreat-
ment reduces the concentration of lignin-derived inhibitors, such as phe-
nolic compounds or furan derivatives. These inhibitors can hinder
enzymatic hydrolysis or microbial fermentation. Enzyme-assisted pre-
treatment helps to remove or modify these inhibitory compounds, leading
to improved conversion rates and yields.
(b) Detoxification of Hemicellulosic Hydrolysates: Enzymatic pretreatment
can also detoxify hemicellulosic hydrolysates by removing or modifying
inhibitory compounds, such as acetic acid, furfural, or hydroxymethylfur-
fural (HMF). These compounds can inhibit enzymatic or microbial activi-
ties and reduce the efficiency of subsequent conversion processes.
Enzyme-assisted pretreatment enhances the detoxification of ­hemicellulosic
hydrolysates, allowing for improved performance during downstream
processes.
(iv) Preservation of Valuable Components
Enzyme-assisted pretreatment aims to selectively modify the recalcitrant
components of biomass waste while preserving valuable components for sub-
sequent conversion processes. The effects of enzyme-assisted pretreatment on
valuable components include the following:
(a) Preservation of Cellulose: Enzymatic pretreatment selectively targets lig-
nin and hemicellulose, while preserving the cellulose component of bio-
mass. This ensures that cellulose, the primary target for enzymatic
hydrolysis, remains intact and accessible for efficient conversion into sug-
ars or other valuable products.
(b) Preservation of Extractable Compounds: Enzyme-assisted pretreatment
can preserve extractable compounds, such as oils, flavors, or bioactive
molecules present in biomass waste. By selectively modifying the biomass
structure, enzymes can facilitate the extraction of valuable compounds
without significant degradation or loss.

5 Enzymatic Hydrolysis of Biomass Waste

Enzymatic hydrolysis is a fundamental step in the valorization of biomass waste,


playing a crucial role in converting complex biomass polymers into valuable prod-
ucts. It involves the use of specific enzymes to degrade cellulose and hemicellulose
into fermentable sugars, which can be further utilized for the production of biofuels,
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 87

biochemicals, or other high-value products (Houfani et al., 2020). The enzymatic


hydrolysis of biomass waste holds significant importance for several reasons.
(i) Conversion of Cellulose into Fermentable Sugars
Cellulose, the main component of biomass waste, is a highly abundant and
renewable polysaccharide. However, its crystalline structure and complex
organization make it resistant to direct enzymatic degradation. Enzymatic
hydrolysis addresses this challenge by utilizing cellulases, including endoglu-
canases, exoglucanases, and β-glucosidases, to break down cellulose into glu-
cose and other fermentable sugars. These sugars serve as valuable substrates
for subsequent fermentation processes, enabling the production of biofuels,
such as ethanol, or biochemicals, such as organic acids or platform chemicals
(Houfani et al., 2020).
(ii) Extraction of Hemicellulosic Sugars
Hemicellulose, another component of biomass waste, consists of a diverse
range of polysaccharides, such as xylan, mannan, and xyloglucan. Enzymatic
hydrolysis of hemicellulose involves the use of specific hemicellulases, such as
xylanases or mannanases, to break down hemicellulosic polymers into their
constituent sugar monomers, such as xylose and mannose. These ­hemicellulosic
sugars are valuable feedstocks for various fermentation processes, contributing
to the production of biofuels, biochemicals, or other bioproducts (Houfani
et al., 2020).
(iii) Efficiency and Specificity of Enzymatic Action
Enzymatic hydrolysis offers high specificity and selectivity in biomass
waste degradation. Enzymes, such as cellulases and hemicellulases, act spe-
cifically on the target polymers, cleaving the glycosidic bonds and releasing
monosaccharides. Unlike chemical or thermochemical methods, enzymatic
hydrolysis avoids the production of unwanted by-products or the degradation
of valuable components. Enzymes can be tailored to the specific composition
of biomass waste, allowing for efficient and selective degradation, leading to
higher yields of desired products (Yang et al., 2011).
(iv) Environmental Sustainability
Enzymatic hydrolysis is an environmentally sustainable approach for bio-
mass waste valorization. Compared to traditional chemical or thermochemical
methods, enzymatic hydrolysis operates under mild conditions, typically at
moderate temperatures and near-neutral pH. This reduces energy consumption,
minimizes the generation of harmful by-products, and contributes to a lower
carbon footprint. Enzymes used in hydrolysis can be derived from renewable
sources or produced through microbial fermentation, further enhancing the
sustainability of the process (Manikandan et al., 2023).
(v) Compatibility with Diverse Biomass Feedstocks
Enzymatic hydrolysis is highly versatile and compatible with a wide range
of biomass feedstocks. It can be applied to lignocellulosic materials, such as
agricultural residues (corn stover, wheat straw), dedicated energy crops
(switchgrass, miscanthus), or forest residues (wood chips, sawdust). Enzymes
88 U. L. Usman et al.

used in hydrolysis can be optimized and tailored to the specific composition


and structure of different biomass sources, ensuring efficient degradation and
high product yields (Saini et al., 2015).
(vi) Process Integration and Valorization of Waste Streams
Enzymatic hydrolysis can be seamlessly integrated with other biomass con-
version processes, such as enzymatic fermentation or downstream purification
steps. The integration allows for the utilization of waste streams or by-products
generated from other processes, enabling a more comprehensive and efficient
valorization of biomass waste (Sharma et al., 2023a). For example, lignin or
other by-products from enzymatic hydrolysis can be further processed and uti-
lized for the production of value-added chemicals or materials.

6 Enzyme-Mediated Conversion of Biomass Waste into


Value-Added Products

Enzyme-mediated conversion of biomass waste offers tremendous potential for the


production of a wide range of value-added products. By harnessing the power of
enzymes, various biomass components can be transformed into renewable fuels,
chemicals, materials, and other high-value products (Bhardwaj et al., 2021). Below
is an overview of the key value-added products derived from biomass waste through
enzyme-mediated conversion.
(i) Biofuels
Biomass waste can be enzymatically converted into biofuels, serving as a
sustainable alternative to fossil fuels. Enzymatic hydrolysis breaks down cel-
lulose and hemicellulose into fermentable sugars, which can be further con-
verted into biofuels through microbial fermentation. The main biofuels derived
from biomass waste include the following:
(a) Bioethanol: Enzymatic hydrolysis followed by fermentation of sugars pro-
duces bioethanol, a renewable fuel with applications in transportation and
as a blend in gasoline.
(b) Biodiesel: Enzymatic conversion of lipids extracted from biomass waste,
such as algae or waste cooking oil, into biodiesel offers a greener alterna-
tive to petroleum-based diesel.
(c) Biogas: Anaerobic digestion of biomass waste, such as agricultural resi-
dues or organic waste, yields biogas rich in methane. Enzymatic hydroly-
sis can enhance the production of fermentable sugars for improved
biogas yield.
(ii) Biochemicals and Platform Chemicals
Biomass waste can be enzymatically converted into a range of bio-based
chemicals, replacing their fossil fuel-derived counterparts (Ewing et al., 2022).
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 89

Enzymatic hydrolysis followed by microbial fermentation enables the produc-


tion of bio-based chemicals and platform chemicals, including:
(a) Organic acids: Enzymatic conversion of sugars derived from biomass
waste can yield organic acids such as lactic acid, acetic acid, or succinic
acid, which find applications in food, pharmaceuticals, and chemical
industries.
(b) Polyols: Biomass waste-derived sugars can be enzymatically converted
into polyols such as xylitol or sorbitol, which have applications in food,
pharmaceuticals, and as renewable building blocks for polymers.
(c) Amino Acids: Enzymatic conversion of biomass waste can yield amino
acids, such as glutamic acid or lysine, which are used in food, animal feed,
and pharmaceutical applications.
(d) Platform Chemicals: Biomass waste can serve as a valuable source of plat-
form chemicals such as levulinic acid, furfural, and ­hydroxymethylfurfural
(HMF), which can be further transformed into a variety of chemicals and
materials.
(iii) Biopolymers and Biomaterials
Biomass waste can be enzymatically converted into biopolymers and bio-
materials, offering sustainable alternatives to petroleum-based plastics and
materials (Narancic et al., 2020). Enzymatic processes enable the conversion
of biomass components into the following:
(a) Bioplastics: Enzymatically derived sugars can be fermented to produce
biopolymers such as polylactic acid (PLA), polyhydroxyalkanoates
(PHA), and cellulose-based materials, which have applications in packag-
ing, textiles, and biomedical industries.
(b) Bio-Based Fibers: Biomass waste-derived cellulose can be enzymatically
processed and spun into bio-based fibers for textile applications, reducing
reliance on synthetic fibers.
(c) Lignin-Based Materials: Enzymatic modification of lignin, a component
of biomass waste, can yield lignin-based materials with applications in
adhesives, composites, or coatings.
(iv) Animal Feed and Fertilizers
Biomass waste can be enzymatically treated to improve its nutritional value
for animal feed or to produce organic fertilizers. Enzymatic processes can
enhance the breakdown of biomass components, making them more digestible
and bioavailable for animal nutrition (Yafetto et al., 2023).
(a) Animal Feed: Enzymatic treatment of biomass waste can improve its
digestibility and nutrient content and remove antinutritional factors, mak-
ing it suitable for incorporation into animal feed formulations.
(b) Fertilizers: Enzymatic conversion of biomass waste can yield organic fer-
tilizers rich in nutrients, providing a sustainable source of plant nutrients
while minimizing environmental impacts.
90 U. L. Usman et al.

7 Challenges and Opportunities in the Enzymatic


Conversion Process

Enzymatic conversion processes offer great potential for the sustainable utilization
of biomass waste. However, they also present various challenges that need to be
addressed for widespread implementation. At the same time, these challenges also
present opportunities for further research and innovation (Benti et al., 2021). Table 4
depicts some key challenges and opportunities in the enzymatic conversion process.

Table 4 Challenges and opportunities in the enzymatic conversion process


Category Challenges Opportunity
Substrate Biomass waste exhibits complex and Developing robust enzyme cocktails or
complexity variable composition, making it engineered enzymes with improved
and variability challenging to develop efficient substrate specificity and tolerance to
enzymatic conversion processes. inhibitors can enhance the efficiency of
The presence of lignin, structural enzymatic conversion. Furthermore,
complexity, and varying ratios of advances in pretreatment technologies
cellulose and hemicellulose affect can help overcome substrate complexity
enzyme accessibility and hydrolysis and improve the accessibility of
efficiency enzymes to biomass components
Enzyme cost Enzymes used in biomass Research efforts to optimize enzyme
and availability conversion processes can be costly, production, reduce costs, and improve
limiting the economic feasibility of enzyme stability can enhance the
enzymatic approaches. Availability commercial viability of enzymatic
and sourcing of enzymes at the conversion processes. Strategies such as
required scale also pose challenges enzyme recycling, immobilization, and
for large-scale implementation the development of robust enzyme
production systems can contribute to
cost reduction and increased availability
Enzyme Enzymes may exhibit reduced Protein engineering techniques can be
stability and stability under the harsh conditions employed to improve enzyme stability
compatibility encountered in biomass conversion and activity under challenging
processes, such as high conditions. Engineering enzymes with
temperatures, extreme pH, or the increased tolerance to temperature, pH,
presence of inhibitors. Compatibility or inhibitors can enhance their
issues between enzymes and other performance in biomass conversion
process conditions can hinder processes. Formulating enzyme cocktails
enzyme performance with complementary stability profiles
can also overcome compatibility issues
and improve overall process efficiency
Process Integrating enzymatic conversion Comprehensive process design and
integration and processes with other process steps, optimization, including reactor design,
scale-up such as pretreatment, fermentation, mass transfer considerations, and process
or downstream processing, can be modeling, can enable efficient integration
challenging. Scaling up enzymatic and scale-up of enzymatic conversion
processes from the laboratory scale processes. Collaboration between
to the industrial scale also presents researchers, industry, and policymakers
technical and economic complexities can facilitate the development of
standardized protocols and guidelines for
scaling up enzymatic processes
(continued)
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 91

Table 4 (continued)
Category Challenges Opportunity
Product Enzymatic conversion processes Optimization strategies, such as enzyme
selectivity and may generate by-products or exhibit engineering, metabolic engineering, and
yield suboptimal selectivity, limiting the process parameter optimization, can
optimization overall yield of the desired products. enhance product selectivity and yield.
Competitive reactions, enzyme Improving enzyme specificity,
kinetics, and metabolic pathways modifying enzyme kinetics, or
can influence product selectivity and implementing process control strategies
yield can help achieve higher conversion
efficiencies and desired product yields
Techno-­ Achieving techno-economic Continual research and development
economic feasibility is crucial for the focused on enzyme cost reduction,
feasibility successful commercialization of process optimization, and valorization of
enzymatic conversion processes. low-cost biomass feedstocks can
The cost of enzymes, biomass enhance the techno-economic feasibility
feedstock, process integration, and of enzymatic conversion processes.
downstream processing must be Integration of enzymatic processes with
balanced to ensure economic other value-added products or
viability by-products can improve overall process
economics

8 Potential Applications of Emerging Technologies


in Enzymatic Valorization

Emerging technologies in enzymatic valorization have the potential to revolutionize


biomass waste management by enabling more efficient and sustainable utilization
of renewable resources. These technologies leverage advancements in enzyme engi-
neering, bioprocessing, and system optimization to expand the range of applications
and enhance the value-added products derived from biomass waste (Wiltschi et al.,
2020). Below are some potential applications of emerging technologies in enzy-
matic valorization.

8.1 Advanced Biofuels Production

Emerging technologies can contribute to the development of advanced biofuels


through the enzymatic valorization of biomass waste. Enzyme engineering and opti-
mization enable efficient conversion of lignocellulosic biomass into biofuels such as
cellulosic ethanol, bio-butanol, and biodiesel. Integrated enzymatic processes, cou-
pled with microbial fermentation or chemical conversion, can enhance the produc-
tion of advanced biofuels with improved energy density and reduced environmental
impact (Sharma et al., 2023a).
92 U. L. Usman et al.

8.2 Specialty Chemicals and Biochemicals

Enzymatic valorization technologies offer the potential for the production of a wide
range of speciality chemicals and biochemicals from biomass waste. By utilizing tai-
lored enzyme systems, enzyme cocktails, and multienzyme platforms, biomass waste
components can be selectively converted into valuable intermediates, such as organic
acids, amino acids, and platform chemicals. These chemicals find applications in the
pharmaceutical, cosmetic, and chemical industries, providing sustainable alternatives
to fossil fuel-derived products (Patel & Shah, 2021).

8.3 Bioplastics and Biopolymers

Enzymatic valorization technologies can contribute to the production of bioplastics


and biopolymers from biomass waste. Through enzymatic hydrolysis and polymer-
ization processes, biomass components such as cellulose, hemicellulose, and lignin
can be converted into biodegradable plastics and polymers. Enzyme engineering
and immobilization techniques enhance the catalytic efficiency and stability of
enzymes, facilitating the synthesis of high-performance biopolymers with tailored
properties (Singhania et al., 2022).

8.4 Nutraceuticals and Functional Ingredients

Emerging enzymatic valorization technologies enable the production of nutraceuti-


cals and functional ingredients from biomass waste. Enzymatic processes can be
employed to extract, modify, or transform bioactive compounds present in biomass
waste into value-added products. For example, enzymes can facilitate the extraction
and modification of antioxidants, flavonoids, or polyphenols for use in dietary sup-
plements, functional foods, or personal care products (Bala et al., 2023).

8.5 Waste Remediation and Environmental Applications

Enzymatic valorization technologies hold potential for waste remediation and envi-
ronmental applications. Enzymes can be utilized to degrade and detoxify pollutants
present in biomass waste or wastewater streams. Enzymatic processes can also be
employed for the bioremediation of contaminated soils or the treatment of industrial
effluents. These applications contribute to the sustainable management of biomass
waste and the mitigation of environmental pollution (Bala et al., 2022).
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 93

8.6 Circular Bioeconomy and Integrated Biorefineries

Emerging enzymatic valorization technologies play a vital role in establishing a


circular bioeconomy by enabling the integration of biomass waste valorization
within integrated biorefineries. Integrated biorefineries leverage multiple processes,
including enzymatic conversion, pretreatment, fermentation, chemical conversion,
and downstream processing, to maximize the utilization of biomass feedstocks and
generate a diverse range of value-added products. Enzymatic valorization technolo-
gies provide efficient and sustainable pathways for the conversion of biomass waste
into multiple products, contributing to the circular utilization of resources (Colussi
et al., 2021).

9 Conclusion and Future Research Prospects

Enzyme-mediated strategies offer a sustainable and efficient pathway for the man-
agement and valorization of biomass waste. This chapter has explored the diverse
and promising field of enzyme-mediated strategies for the effective management
and valorization of biomass waste. Through an in-depth examination of various
aspects, we have gained valuable insights into the role of enzymes in biomass waste
processing and the potential they hold for sustainable resource utilization. Biomass
waste poses significant environmental challenges, including waste accumulation,
resource depletion, and greenhouse gas emissions. However, through effective man-
agement and valorization strategies, biomass waste can be transformed into valu-
able products, mitigating environmental impact and contributing to a circular
bioeconomy. Enzymes play a central role in biomass waste processing, offering
unique capabilities for the degradation, pretreatment, hydrolysis, and conversion of
biomass components. This chapter has provided a comprehensive understanding of
the types of biomass waste suitable for enzymatic degradation, the major enzymes
involved, and the mechanisms and factors influencing enzyme activity. We have also
explored recent advancements in enzyme engineering, which hold promise for
enhancing degradation efficiency and expanding enzyme functionality. Enzymatic
pretreatment and hydrolysis have emerged as crucial steps in biomass waste valori-
zation. By breaking down complex biomass structures and converting them into
more accessible forms, enzymes facilitate the efficient utilization of biomass waste.
This chapter has elucidated the various enzymatic pretreatment methods, the types
of enzymes used for hydrolysis, and the factors influencing enzymatic hydrolysis
efficiency. Furthermore, the enzyme-mediated conversion of biomass waste into
value-added products has been discussed extensively. By harnessing the versatility
of enzymes and their ability to target specific biomass components, various value-­
added products such as biofuels, speciality chemicals, bioplastics, and nutraceuti-
cals can be produced.
94 U. L. Usman et al.

Future research directions include the following:


(i) Discovery and Engineering of Novel Enzymes
Future research should focus on exploring and harnessing novel enzymes
from diverse sources, such as extremophiles, metagenomics, and genetically
modified microorganisms. By identifying enzymes with unique properties and
high substrate specificity, researchers can expand the enzymatic toolbox for
biomass waste valorization. Additionally, enzyme engineering techniques,
including protein design and directed evolution, can be employed to enhance
enzyme performance and tailor their activity to specific waste streams.
(ii) Optimization of Enzyme Production and Utilization
Efficient and cost-effective enzyme production is critical for large-scale
implementation. Future research should investigate strategies to improve
enzyme yield, stability, and compatibility with different substrates and operat-
ing conditions. This includes exploring alternative enzyme production hosts,
developing innovative fermentation processes, and optimizing enzyme immo-
bilization techniques. Furthermore, the integration of in situ enzyme produc-
tion within biorefinery systems can enhance process efficiency and reduce costs.
(iii) Exploration of Advanced Biorefinery Concepts
Biorefineries play a vital role in converting biomass waste into value-added
products. Future research should focus on exploring advanced biorefinery con-
cepts, such as hybrid processes, cascading utilization of biomass components,
and integrated biorefinery systems. Enzyme-mediated strategies can be opti-
mized and integrated with other conversion technologies, such as microbial
fermentation, thermochemical conversion, and electrochemical processes.
These interdisciplinary approaches will enable the efficient utilization of
diverse biomass feedstocks and maximize the production of fuels, chemicals,
materials, and energy.
(iv) Life-Cycle Assessment and Sustainability Analysis
To ensure the sustainability of enzyme-mediated strategies, future research
should include a comprehensive life-cycle assessment (LCA) and sustainabil-
ity analysis. This involves evaluating the environmental, economic, and social
impacts of enzyme-mediated processes throughout their life cycle. LCA can
guide process optimization, identify hotspots, and inform decision-making to
minimize environmental footprints and promote sustainable waste manage-
ment practices.

References

Abdel-Shafy, H. I., & Mansour, M. S. M. (2018). Solid waste issue: Sources, composition, dis-
posal, recycling, and valorization. Egyptian Journal of Petroleum. https://doi.org/10.1016/j.
ejpe.2018.07.003
Abraham, A., Mathew, A. K., Park, H., Choi, O., Sindhu, R., Parameswaran, B., Pandey, A., Park,
J. H., & Sang, B. I. (2020). Pretreatment strategies for enhanced biogas production from lig-
nocellulosic biomass. Bioresource Technology. https://doi.org/10.1016/j.biortech.2019.122725
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 95

Amalina, F., Razak, A. S. A., Krishnan, S., Sulaiman, H., Zularisam, A. W., & Nasrullah, M. (2022).
Biochar production techniques utilizing biomass waste-derived materials and environmental
applications – A review. Journal of Hazardous Materials Advances. https://doi.org/10.1016/j.
hazadv.2022.100134
Awogbemi, O., & Von Kallon, D. V. (2022). Valorization of agricultural wastes for biofuel applica-
tions. Heliyon. https://doi.org/10.1016/j.heliyon.2022.e11117
Bala, S., Garg, D., Thirumalesh, B. V., Sharma, M., Sridhar, K., Inbaraj, B. S., & Tripathi,
M. (2022). Recent strategies for bioremediation of emerging pollutants: A review for a green
and sustainable environment. Toxics. https://doi.org/10.3390/toxics10080484
Bala, S., Garg, D., Sridhar, K., Inbaraj, B. S., Singh, R., Kamma, S., Tripathi, M., & Sharma,
M. (2023). Transformation of agro-waste into value-added bioproducts and bioactive
compounds: Micro/nano formulations and application in the agri-food-pharma sector.
Bioengineering. https://doi.org/10.3390/bioengineering10020152
Balan, V. (2014). Current challenges in commercially producing biofuels from lignocellulosic bio-
mass. ISRN Biotechnology. https://doi.org/10.1155/2014/463074
Benti, N. E., Gurmesa, G. S., Argaw, T., Aneseyee, A. B., Gunta, S., Kassahun, G. B., Aga, G. S.,
& Asfaw, A. A. (2021). The current status, challenges and prospects of using biomass energy in
Ethiopia. Biotechnology for Biofuels. https://doi.org/10.1186/s13068-­021-­02060-­3
Bhardwaj, N., Kumar, B., Agrawal, K., & Verma, P. (2021). Current perspective on production and
applications of microbial cellulases: A review. Bioresources and Bioprocessing. https://doi.
org/10.1186/s40643-­021-­00447-­6
Chen, C. C., Dai, L., Ma, L., & Guo, R. T. (2020). Enzymatic degradation of plant biomass and
synthetic polymers. Nature Reviews Chemistry. https://doi.org/10.1038/s41570-­020-­0163-­6
Colussi, F., Michelin, M., Gomes, D. G., Rocha, C. M. R., Romaní, A., Domingues, L., & Teixeira,
J. A. (2021). Integrated technologies for extractives recovery, fractionation, and bioethanol produc-
tion from lignocellulose. Biomass, Biofuels, Biochemicals: Circular Bioeconomy: Technologies
for Biofuels and Biochemicals. https://doi.org/10.1016/B978-­0-­323-­89855-3.00001-­7
Estela, R., & Luis, J. (2013). Hydrolysis of biomass mediated by cellulases for the production of
sugars. Sustainable Degradation of Lignocellulosic Biomass – Techniques, Applications and
Commercialization. https://doi.org/10.5772/53719
Ewing, T. A., Nouse, N., van Lint, M., van Haveren, J., Hugenholtz, J., & van Es, D. S. (2022).
Fermentation for the production of biobased chemicals in a circular economy: A perspective
for the period 2022–2050. Green Chemistry. https://doi.org/10.1039/d1gc04758b
Fülöp, L., & Ecker, J. (2020). An overview of biomass conversion: Exploring new opportunities.
PeerJ. https://doi.org/10.7717/peerj.9586
German, D. P., Chacon, S. S., & Allison, S. D. (2011). Substrate concentration and enzyme alloca-
tion can affect rates of microbial decomposition. Ecology. https://doi.org/10.1890/10-­2028.1
Guo, H., Zhao, Y., Chang, J. S., & Lee, D. J. (2023). Enzymes and enzymatic mechanisms in enzy-
matic degradation of lignocellulosic biomass: A mini-review. Bioresource Technology. https://
doi.org/10.1016/j.biortech.2022.128252
Homaei, A. A., Sariri, R., Vianello, F., & Stevanato, R. (2013). Enzyme immobilization: An update.
Journal of Chemical Biology. https://doi.org/10.1007/s12154-­013-­0102-­9
Houfani, A. A., Anders, N., Spiess, A. C., Baldrian, P., & Benallaoua, S. (2020). Insights from
enzymatic degradation of cellulose and hemicellulose to fermentable sugars– A review.
Biomass and Bioenergy. https://doi.org/10.1016/j.biombioe.2020.105481
Jayasekara, S., & Ratnayake, R. (2019). Microbial cellulases: An overview and applications.
Cellulose. https://doi.org/10.5772/intechopen.84531
Jekayinfa, S. O., Orisaleye, J. I., & Pecenka, R. (2020). An assessment of potential resources for
biomass energy in Nigeria. Resources. https://doi.org/10.3390/resources9080092
Kumar, A., & Chandra, R. (2020). Ligninolytic enzymes and its mechanisms for degradation of
lignocellulosic waste in environment. Heliyon. https://doi.org/10.1016/j.heliyon.2020.e03170
Manikandan, S., Vickram, S., Sirohi, R., Subbaiya, R., Krishnan, R. Y., Karmegam, N.,
Sumathijones, C., Rajagopal, R., Chang, S. W., Ravindran, B., & Awasthi, M. K. (2023).
Critical review of biochemical pathways to transformation of waste and biomass into bioen-
ergy. Bioresource Technology. https://doi.org/10.1016/j.biortech.2023.128679
96 U. L. Usman et al.

Mishra, S., Singh, P. K., Dash, S., & Pattnaik, R. (2018). Microbial pretreatment of lignocel-
lulosic biomass for enhanced biomethanation and waste management. 3 Biotech. https://doi.
org/10.1007/s13205-­018-­1480-­z
Mukherjee, J., & Gupta, M. N. (2015). Biocatalysis for biomass valorization. Sustainable Chemical
Processes. https://doi.org/10.1186/s40508-­015-­0037-­2
Narancic, T., Cerrone, F., Beagan, N., & O’Connor, K. E. (2020). Recent advances in bioplastics:
Application and biodegradation. Polymers. https://doi.org/10.3390/POLYM12040920
Nargotra, P., Sharma, V., Lee, Y. C., Tsai, Y. H., Liu, Y. C., Shieh, C. J., Tsai, M. L., Di Dong, C.,
& Kuo, C. H. (2023). Microbial lignocellulolytic enzymes for the effective valorization of
Lignocellulosic biomass: A review. Catalysts. https://doi.org/10.3390/catal13010083
Nizami, A. S., Rehan, M., Waqas, M., Naqvi, M., Ouda, O. K. M., Shahzad, K., Miandad, R.,
Khan, M. Z., Syamsiro, M., Ismail, I. M. I., & Pant, D. (2017). Waste biorefineries: Enabling
circular economies in developing countries. Bioresource Technology. https://doi.org/10.1016/j.
biortech.2017.05.097
Ogbu, C. C., & Okechukwu, S. N. (2023). Agro-industrial waste management: The circular and
bioeconomic perspective. Agricultural Waste – New Insights [Working Title]. https://doi.
org/10.5772/intechopen.109181
Okafor, C. C., Nzekwe, C. A., Ajaero, C. C., Ibekwe, J. C., & Otunomo, F. A. (2022). Biomass
utilization for energy production in Nigeria: A review. Cleaner Energy Systems. https://doi.
org/10.1016/j.cles.2022.100043
Patel, A., & Shah, A. R. (2021). Integrated lignocellulosic biorefinery: Gateway for production of
second generation ethanol and value added products. Journal of Bioresources and Bioproducts.
https://doi.org/10.1016/j.jobab.2021.02.001
Ramos, L. P., Suota, M. J., Fockink, D. H., Pavaneli, G., Da Silva, T. A., & Łukasik, R. M. (2020).
Enzymes and biomass pretreatment. Recent Advances in Bioconversion of Lignocellulose to
Biofuels and Value Added Chemicals Within the Biorefinery Concept. https://doi.org/10.1016/
B978-­0-­12-­818223-­9.00004-­7
Robinson, P. K. (2015). Enzymes: Principles and biotechnological applications. Essays in
Biochemistry. https://doi.org/10.1042/BSE0590001
Saini, J. K., Saini, R., & Tewari, L. (2015). Lignocellulosic agriculture wastes as biomass feed-
stocks for second-generation bioethanol production: Concepts and recent developments. 3
Biotech. https://doi.org/10.1007/s13205-­014-­0246-­5
Sakhuja, D., Ghai, H., Rathour, R. K., Kumar, P., Bhatt, A. K., & Bhatia, R. K. (2021). Cost-­
effective production of biocatalysts using inexpensive plant biomass: A review. 3 Biotech.
https://doi.org/10.1007/s13205-­021-­02847-­z
Sasmoko, Zaman, K., Malik, M., Awan, U., Handayani, W., Jabor, M. K., & Asif, M. (2022).
Environmental effects of bio-waste recycling on industrial circular economy and eco-­
sustainability. Recycling. https://doi.org/10.3390/recycling7040060
Sharma, L., Alam, N. M., Roy, S., Satya, P., Kar, G., Ghosh, S., Goswami, T., & Majumdar,
B. (2023a). Optimization of alkali pretreatment and enzymatic saccharification of jute
(Corchorus olitorius L.) biomass using response surface methodology. Bioresource Technology.
https://doi.org/10.1016/j.biortech.2022.128318
Sharma, S., Tsai, M. L., Sharma, V., Sun, P. P., Nargotra, P., Bajaj, B. K., Chen, C. W., & Dong,
C. D. (2023b). Environment friendly pretreatment approaches for the bioconversion of
Lignocellulosic biomass into biofuels and value-added products. Environments – MDPI. https://
doi.org/10.3390/environments10010006
Singh, R. K., Tiwari, M. K., Singh, R., & Lee, J. K. (2013). From protein engineering to immobi-
lization: Promising strategies for the upgrade of industrial enzymes. International Journal of
Molecular Sciences. https://doi.org/10.3390/ijms14011232
Singh, R., Langyan, S., Rohtagi, B., Darjee, S., Khandelwal, A., Shrivastava, M., Kothari, R.,
Mohan, H., Raina, S., Kaur, J., & Singh, A. (2022). Production of biofuels options by contribu-
tion of effective and suitable enzymes: Technological developments and challenges. Materials
Science for Energy Technologies. https://doi.org/10.1016/j.mset.2022.05.001
Enzyme-Mediated Strategies for Effective Management and Valorization of Biomass… 97

Singhania, R. R., Patel, A. K., Raj, T., Chen, C. W., Ponnusamy, V. K., Tahir, N., Kim, S. H., &
Dong, C. D. (2022). Lignin valorisation via enzymes: A sustainable approach. Fuel. https://doi.
org/10.1016/j.fuel.2021.122608
Song, B., Lin, R., Lam, C. H., Wu, H., Tsui, T. H., & Yu, Y. (2021). Recent advances and challenges
of inter-disciplinary biomass valorization by integrating hydrothermal and biological tech-
niques. Renewable and Sustainable Energy Reviews. https://doi.org/10.1016/j.rser.2020.110370
Sweeney, M. D., & Xu, F. (2012). Biomass converting enzymes as industrial biocatalysts for fuels
and chemicals: Recent developments. Catalysts. https://doi.org/10.3390/catal2020244
Wesseler, J., Kleter, G., Meulenbroek, M., & Purnhagen, K. P. (2023). EU regulation of genetically
modified microorganisms in light of new policy developments: Possible implications for EU
bioeconomy investments. Applied Economic Perspectives and Policy. https://doi.org/10.1002/
aepp.13259
Wiltschi, B., Cernava, T., Dennig, A., Galindo Casas, M., Geier, M., Gruber, S., Haberbauer, M.,
Heidinger, P., Herrero Acero, E., Kratzer, R., Luley-Goedl, C., Müller, C. A., Pitzer, J., Ribitsch,
D., Sauer, M., Schmölzer, K., Schnitzhofer, W., Sensen, C. W., Soh, J., et al. (2020). Enzymes
revolutionize the bioproduction of value-added compounds: From enzyme discovery to special
applications. Biotechnology Advances. https://doi.org/10.1016/j.biotechadv.2020.107520
Yaashikaa, P. R., Devi, M. K., & Kumar, P. S. (2022). Advances in the application of immobi-
lized enzyme for the remediation of hazardous pollutant: A review. Chemosphere. https://doi.
org/10.1016/j.chemosphere.2022.134390
Yafetto, L., Odamtten, G. T., & Wiafe-Kwagyan, M. (2023). Valorization of agro-industrial wastes
into animal feed through microbial fermentation: A review of the global and Ghanaian case.
Heliyon. https://doi.org/10.1016/j.heliyon.2023.e14814
Yang, B., Dai, Z., Ding, S. Y., & Wyman, C. E. (2011). Enzymatic hydrolysis of cellulosic bio-
mass. Biofuels. https://doi.org/10.4155/bfs.11.116
Zamora Zamora, H. D., de Freitas, C., Bueno, D., Shimizu, F. L., Contiero, J., & Brienzo,
M. (2020). Biomass Fractionation Based on Enzymatic Hydrolysis for Biorefinery Systems.
https://doi.org/10.1007/978-­981-­15-­9593-­6_9
Zhao, B., Al Rasheed, H., Ali, I., & Hu, S. (2021). Efficient enzymatic saccharification of alkaline
and ionic liquid-pretreated bamboo by highly active extremozymes produced by the co-culture of
two halophilic fungi. Bioresource Technology. https://doi.org/10.1016/j.biortech.2020.124115
Zhou, C., & Wang, Y. (2020). Recent progress in the conversion of biomass wastes into functional
materials for value-added applications. Science and Technology of Advanced Materials. https://
doi.org/10.1080/14686996.2020.1848213
Zhu, B., Wang, D., & Wei, N. (2022). Enzyme discovery and engineering for sustainable plastic
recycling. Trends in Biotechnology. https://doi.org/10.1016/j.tibtech.2021.02.008
Nanotechnological Advancements
for Enhancing Lignocellulosic Biomass
Valorization

Vijayalakshmi Ghosh

1 Introduction

The world is currently facing two major challenges, the need to transition to renew-
able energy sources and the imperative to adopt sustainable materials. These chal-
lenges are driven by concerns over climate change, finite fossil fuel reserves, and
the detrimental environmental impact of traditional energy production and material
extraction. The world’s dependence on fossil fuels, such as coal, oil, and natural gas,
has resulted in significant greenhouse gas emissions, contributing to climate change
and global warming. Renewable energy sources, including solar, wind, hydroelec-
tric, geothermal, and biomass, offer a promising solution to mitigate these environ-
mental issues. The growing global demand for renewable energy is fueled by the
recognition that a transition to sustainable alternatives is essential to reduce carbon
emissions and combat the adverse effects of climate change. The pressing need to
combat climate change and reduce our dependence on finite fossil fuels has driven
the global demand for renewable energy sources and sustainable materials.
As the world strives to transition toward a greener and more sustainable future,
lignocellulosic biomass valorization has emerged as a promising solution.
Lignocellulosic biomass, derived from plant materials such as agricultural residues,
forest waste, and dedicated energy crops, holds immense potential as a renewable
feedstock for the production of bioenergy, biofuels, and sustainable materials (Chen,
2017; Rai et al., 2022). The global demand for renewable energy sources and sus-
tainable materials has sparked a growing interest in lignocellulosic biomass valori-
zation. The depletion of fossil fuel reserves and the harmful environmental impact
of their extraction and combustion have led to an urgent quest for alternative energy
sources. Renewable energy, such as solar, wind, hydroelectric, and geothermal
power, has gained traction due to its low-carbon footprint and potential to reduce

V. Ghosh (*)
GD Rungta College of Science & Technology, Bhilai, Chhattisgarh, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 99


A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_5
100 V. Ghosh

greenhouse gas emissions. However, these sources often suffer from intermittent
availability and energy storage challenges. Lignocellulosic biomass valorization
offers a complementary solution by providing a storable and dispatchable source of
renewable energy (Elumalai et al., 2018; Scarlat et al., 2015; Valdivia et al., 2016).
Lignocellulosic biomass is abundant and widely available, making it an attractive
feedstock for various applications. Its versatility allows it to be converted into mul-
tiple valuable products, including biofuels, bioproducts, and sustainable materials.
This multifunctionality of lignocellulosic biomass makes it a significant contender
in meeting the diverse demands of a renewable and sustainable future. Bioenergy,
derived from the conversion of biomass, plays a pivotal role in mitigating green-
house gas emissions and enhancing energy security. Lignocellulosic biomass can be
transformed into biofuels such as bioethanol, biobutanol, and biodiesel through bio-
chemical and thermochemical processes. Advanced technologies, such as enzy-
matic hydrolysis and gasification, have improved the efficiency and cost-effectiveness
of these conversion pathways. As a result, lignocellulosic biomass valorization con-
tributes to the reduction of carbon emissions and reliance on fossil fuels. In addition
to bioenergy and biofuels, lignocellulosic biomass valorization offers the potential
for sustainable materials. Biocomposites, bioplastics, and biodegradable packaging
materials can be derived from lignocellulosic sources, reducing the environmental
impact of traditional materials like plastics and concrete. Sustainable materials
made from lignocellulosic biomass promote the principles of circular economies,
where waste is minimized, and resources are used efficiently. The growing interest
in lignocellulosic biomass valorization has been accelerated by significant techno-
logical advancements. Research and development efforts have led to improved pre-
treatment techniques, novel enzyme systems, and innovative conversion processes.
These developments have enhanced the efficiency of lignocellulosic biomass con-
version, making it a more economically viable and scalable option for meeting
global energy and material demands. The increasing awareness of environmental
challenges and the global shift toward sustainable practices have prompted govern-
ments and industries to invest in renewable energy sources and sustainable materi-
als. Supportive policies, subsidies, and incentives have further stimulated interest in
lignocellulosic biomass valorization projects, driving innovation and commercial-
ization efforts worldwide. The global demand for renewable energy sources and
sustainable materials has catalyzed a growing interest in lignocellulosic biomass
valorization. As the world seeks to reduce carbon emissions, combat climate change,
and transition to a more sustainable future, lignocellulosic biomass offers a versatile
and abundant resource with the potential to provide renewable energy, biofuels, and
sustainable materials. Technological advancements and supportive policies have
accelerated the development and commercialization of lignocellulosic biomass con-
version technologies, making them increasingly viable and attractive options for
addressing the pressing challenges of our time. The integration of lignocellulosic
biomass valorization into our energy and material supply chains brings us closer to
a greener, more sustainable, and environmentally conscious world.
Pretreatment of biomass is a significant step required for the removal of lignin
that helps in the modification of the structure and composition of the biomass, thus
Nanotechnological Advancements for Enhancing Lignocellulosic Biomass Valorization 101

allowing efficient hydrolysis in the subsequent step for the production of ferment-
able sugars. The sugars obtained after hydrolysis with the use of suitable microbes
can be employed for the production of biofuels and green chemicals.
Nanotechnology represents a techno-scientific revolution that is being exten-
sively applied in different fields including bioenergy, biofuel, and environmental
management. This chapter brings about an insight into the application of nanobio-
technology in lignocellulosic biomass valorization (Sankaran, 2021).

2 Lignocellulosics: A Potential Source


for Biomass Valorization

In the quest for sustainable and renewable resources, lignocellulosics have emerged
as a promising feedstock for biomass valorization. Lignocellulosic materials, abun-
dant in plant cell walls, are composed of cellulose (38–50%), hemicellulose
(17–32%), and lignin (15–30%), making them a valuable resource with diverse
applications (Garlapati, 2020). Biomass valorization refers to the conversion of lig-
nocellulosic biomass into various high-value products, including bioenergy, biofu-
els, bioproducts, and sustainable materials. Lignocellulosics have the potential as a
versatile source for biomass valorization, emphasizing its significance in fostering a
greener and more sustainable future (Bhatia, 2020; Beisl et al., 2017; Amthor,
2003). Lignocellulosic biomass is readily available and abundant, making it an
attractive alternative to finite fossil fuels and nonrenewable resources. This feed-
stock can be sourced from various agricultural residues, forest waste, dedicated
energy crops, and industrial by-products. As renewable materials, lignocellulosics
offer a sustainable and environmentally friendly solution to meet the increasing
global demand for energy and materials. One of the primary applications of ligno-
cellulosic biomass valorization is bioenergy production. Through biochemical and
thermochemical processes, lignocellulosics can be converted into biofuels like bio-
ethanol, biobutanol, and biohydrogen. The bioenergy derived from lignocellulosics
is a clean and low-carbon alternative to fossil fuels, contributing to greenhouse gas
emissions reduction and climate change mitigation. The versatility of lignocellu-
losic biomass as a bioenergy source helps promote energy security and mitigate the
environmental impact of traditional energy production. Lignocellulosics have the
potential to revolutionize the transportation sector by serving as feedstock for
advanced biofuels. These biofuels, such as cellulosic ethanol and renewable diesel,
can be used in existing internal combustion engines with minimal modifications,
reducing the carbon footprint of transportation significantly. By providing sustain-
able alternatives to conventional fossil fuels, lignocellulosic-based biofuels contrib-
ute to reducing air pollution, dependency on imported oil, and greenhouse gas
emissions from the transportation sector. In addition to bioenergy and biofuels, lig-
nocellulosic biomass valorization can lead to the production of various bioproducts
and sustainable materials. Biocomposites, biodegradable plastics, and biomaterials
102 V. Ghosh

derived from lignocellulosics offer eco-friendly alternatives to conventional materi-


als such as plastics and concrete. These sustainable materials promote circular econ-
omies by reducing waste and extending the lifespan of resources, contributing to a
more resource-efficient and waste-free society. While lignocellulosics offer
immense potential for biomass valorization, their complex and recalcitrant nature
presents challenges. The rigid structure of plant cell walls, primarily composed of
cellulose and hemicellulose surrounded by lignin, hinders efficient enzymatic
hydrolysis and conversion processes. To overcome these challenges, ongoing
research and technological advancements have focused on developing efficient pre-
treatment methods, novel enzyme systems, and innovative conversion processes.
These advancements are paving the way for cost-effective and commercially viable
lignocellulosic biomass valorization.
Lignocellulosics represent a remarkable and versatile source for biomass valori-
zation. With their abundance, renewability, and diverse composition, these materials
offer a sustainable solution to meet the global demand for bioenergy, biofuels, bio-
products, and sustainable materials. The valorization of lignocellulosic biomass
contributes to reducing greenhouse gas emissions, mitigating climate change, and
fostering a greener and more sustainable future. As technology continues to prog-
ress, lignocellulosic biomass valorization will play an increasingly pivotal role in
driving the transition toward a more resource-efficient, eco-friendly, and sustainable
society. Embracing the potential of lignocellulosics is a critical step toward achiev-
ing a more sustainable and prosperous future for generations to come (Jönsson
et al., 2013).

3 Lignocellulosic Biomass: Source and Composition

Lignocellulosic biomass is a valuable and renewable resource derived from plant


materials, which are rich in cellulose, hemicellulose, and lignin. This composition
makes lignocellulosic biomass an essential feedstock for various applications, includ-
ing bioenergy production, biofuels, bioproducts, and sustainable materials.
Lignocellulosic biomass acts as a suitable candidate for sustainable renewable
resources because being the majorly unutilized and are abundantly available world-
wide. Understanding the source and composition of lignocellulosic biomass is crucial
for harnessing its potential to foster a greener and more sustainable future (Fig. 1).

3.1 Sources of Lignocellulosic Biomass

Lignocellulosic biomass can be derived from a diverse range of plant materials,


offering a plentiful and renewable source for various applications (Adewuyi, 2022;
Costa et al., 2015; Kaparaju et al., 2008; Lo et al., 2021; Saini et al., 2015). Some of
the primary sources of lignocellulosic biomass include agricultural residue, forest
Nanotechnological Advancements for Enhancing Lignocellulosic Biomass Valorization 103

Lignocellulose Lignocellulose
Sources Composition

Agricultural
Cellulose
residues

Forest waste Hemicellulose

Energy crops Lignin

Industrial
by-products

Fig. 1 Sources and composition of lignocellulose

waste, energy crops, and industrial by-products. Crop residues such as straw, corn
stover, rice husks, and sugarcane bagasse are abundant sources of lignocellulosic
biomass. These residues are by-products of agricultural activities and represent an
underutilized resource that can be harnessed for sustainable applications. Wood
waste, sawdust, and bark from forestry operations are considered lignocellulosic
biomass sources. As sustainable forest management practices are adopted, the
potential for utilizing forest waste as a valuable feedstock increases. Dedicated
energy crops, such as switchgrass, miscanthus, and woody biomass crops, are grown
specifically for their potential in bioenergy production. These crops offer high yields
and fast growth rates, making them suitable for large-scale biomass cultivation.
Various industries, including the paper and pulp industry, generate significant
amounts of lignocellulosic waste, such as black liquor and paper sludge. Utilizing
these by-products can contribute to waste reduction and the circular economy
(Rezania, 2020).

3.2 Composition of Lignocellulosic Biomass

Lignocellulosic biomass is primarily composed of three main components: cellu-


lose, hemicellulose, and lignin. Each of these components plays a distinct role in the
structure and properties of the plant material. Cellulose is the most abundant com-
ponent of lignocellulosic biomass, constituting about 40–50% of its composition. It
is a linear polysaccharide made up of glucose units linked together, forming long
chains. Cellulose provides structural strength and rigidity to plant cell walls,
104 V. Ghosh

making it a crucial component for the structural integrity of plants. Hemicellulose is


a heterogeneous group of polysaccharides that makes up around 20–35% of the
lignocellulosic biomass. Unlike cellulose, hemicellulose is branched and contains
various sugar units, such as xylose, mannose, and glucose. Hemicellulose contrib-
utes to the flexibility of plant cell walls and plays a significant role in facilitating the
conversion of lignocellulosic biomass into valuable products. Lignin is a complex
aromatic polymer that accounts for about 15–30% of the composition of lignocel-
lulosic biomass. It provides rigidity and strength to plant cell walls, making the
biomass resistant to microbial degradation. Lignin is the primary barrier in the effi-
cient conversion of lignocellulosic biomass into bioenergy and other products, as its
recalcitrance hinders access to cellulose and hemicellulose.
Apart from cellulose, hemicellulose, and lignin, lignocellulosic biomass also
contains minor components such as extractives (e.g., resins, waxes), ash, and other
minor sugars. These components can vary depending on the source of the lignocel-
lulosic biomass and its pretreatment history. Lignocellulosic biomass is a valuable
and renewable resource derived from various plant materials. Its composition, pri-
marily comprising cellulose, hemicellulose, and lignin, makes it a significant feed-
stock for bioenergy production, biofuels, bioproducts, and sustainable materials.
The diversity of lignocellulosic biomass sources, from agricultural residues to forest
waste and energy crops, ensures a reliable and consistent supply for various applica-
tions. Understanding the composition of lignocellulosic biomass and its sources is
essential for harnessing its full potential and advancing a greener and more sustain-
able future. Embracing lignocellulosic biomass as a valuable resource contributes to
reducing carbon emissions, mitigating climate change, and fostering a more envi-
ronmentally conscious and resource-efficient society.

4 Need for Lignocellulosic Biomass Pretreatment

Lignocellulosic biomass, composed of cellulose, hemicellulose, and lignin, is a


promising and renewable resource with diverse applications in bioenergy produc-
tion, biofuels, bioproducts, and sustainable materials. However, the efficient utiliza-
tion of lignocellulosic biomass presents significant challenges due to its complex
and recalcitrant nature. Lignocellulosic biomass pretreatment is a crucial step in the
conversion process that aims to overcome these challenges. The recalcitrance of
lignocellulosic biomass is a major obstacle in its efficient conversion to valuable
products. The rigid and compact structure of plant cell walls, primarily composed of
cellulose and hemicellulose surrounded by lignin, restricts the accessibility of
enzymes and microorganisms to the polysaccharides. As a result, the enzymatic
hydrolysis of cellulose and hemicellulose into fermentable sugars becomes slow
and inefficient, hindering the overall bioconversion process. Lignocellulosic bio-
mass pretreatment is necessary to disrupt this recalcitrant structure, making the cel-
lulose and hemicellulose more accessible for subsequent conversion steps. During
pretreatment, lignin and hemicellulose are partially removed or modified, exposing
cellulose fibers and creating pores in the biomass structure. This enhanced surface
Nanotechnological Advancements for Enhancing Lignocellulosic Biomass Valorization 105

area and improved porosity facilitate the penetration of enzymes, such as cellulases
and hemicellulases, leading to increased enzymatic digestibility. As a result, the
hydrolysis of cellulose and hemicellulose into fermentable sugars is accelerated,
improving the overall efficiency of bioconversion processes. Lignocellulosic bio-
mass contains various inhibitory compounds, such as furans, phenols, and organic
acids, which are generated during biomass pretreatment and hydrolysis. These com-
pounds can interfere with enzyme activity, hinder microbial growth, and negatively
impact fermentation processes. Pretreatment helps to reduce the formation of inhib-
itory compounds and remove or detoxify those already present in the biomass. This
allows for a more favorable environment for enzymatic hydrolysis and microbial
fermentation, leading to higher product yields and improved process economics.
Lignocellulosic biomass pretreatment also aids in the separation of lignin from cel-
lulose and hemicellulose, simplifying downstream processing steps. By partially
delignifying the biomass, pretreatment reduces the load on downstream separation
and purification processes. This process results in the production of cleaner and
more refined streams of cellulose and hemicellulose, which can be directly utilized
in subsequent conversion steps. The success of lignocellulosic biomass conversion
processes is heavily dependent on the efficiency and economics of the overall sys-
tem. By enhancing the accessibility and digestibility of cellulose and hemicellulose,
lignocellulosic biomass pretreatment significantly improves process efficiency,
leading to higher yields of bioenergy, biofuels, and bioproducts. Moreover, pretreat-
ment reduces the enzyme loading required for hydrolysis and allows for faster reac-
tion rates, leading to cost savings and improved commercial viability.
The need for lignocellulosic biomass pretreatment is evident in its ability to over-
come biomass recalcitrance, enhance enzymatic digestibility, reduce inhibitory
compounds, and optimize process efficiency and economics. By breaking down the
complex and rigid structure of lignocellulosic biomass, pretreatment significantly
improves the accessibility of cellulose and hemicellulose, leading to higher yields
of bioenergy, biofuels, and bioproducts (Dahadha et al., 2017).

5 Common Pretreatment Techniques

Several pretreatment techniques have been developed to address the challenges


associated with lignocellulosic biomass conversion (Fig. 2). Some of the most com-
monly employed pretreatment methods include the following: (1) Physical

Physical Chemical Biological


Pre-treatments Pre-treatment Pre-treatments
• Miling • Acids • Microbes
• Grinding • Alkali • Enzymes
• Size reduction • Organic solvents

Fig. 2 Pretreatment methods for lignocellulosic biomass valorization


106 V. Ghosh

pretreatment, where mechanical processes, such as milling, grinding, and size


reduction, can disrupt the biomass structure and increase surface area, aiding in
enzymatic access, (2) chemical pretreatment, where acidic, alkaline, or organic sol-
vents can be used to break down lignin and hemicellulose, making cellulose more
accessible to enzymes, (3) biological pretreatment, where some microorganisms,
such as white-­rot fungi, produce enzymes that selectively degrade lignin, facilitat-
ing biomass delignification, and (4) physicochemical pretreatment, where combin-
ing physical and chemical methods can synergistically enhance biomass accessibility
and enzymatic digestibility.
Various pretreatment techniques, such as physical, chemical, biological, and
physicochemical methods, offer flexible approaches to address the unique charac-
teristics of different biomass sources. Out of these methods, chemical and physico-
chemical are the most efficient pretreatment methods. However, the crux of such
methods is the large expenditure toward heat, power, and manpower with the addi-
tion of environmental issues due to the production of toxic inhibitors. In recent
years, to overcome such problems, nanotechnology-based treatment strategies seem
to be a promising alternative approach that can be integrated into the present pre-
treatment processes. Nanotechnology has emerged as a promising approach to over-
come these challenges, enabling enhanced pretreatment and enzymatic hydrolysis,
thereby accelerating the lignocellulosic biomass valorization process.

6 Nanotechnology in Lignocellulosic Biomass Valorization

Nanobiotechnology, an emerging interdisciplinary field, plays a pivotal role in


enhancing biomass valorization processes. Nanobiotechnological-based biomass
valorization involves the integration of nanotechnology and biotechnology
to improve the efficiency and effectiveness of biomass conversion.
Nanobiotechnological-based biomass valorization represents a significant advance-
ment in the conversion of renewable biomass resources into valuable products.
Nanomaterials, enzymes, microorganisms, research institutions, and commercial
enterprises are key players in driving this field forward. Nanomaterials have revolu-
tionized biomass pretreatment, enabling more efficient enzymatic hydrolysis and
reducing energy consumption. Enhanced enzymes and nanozymes have improved
the efficiency of biomass conversion, while nanocarriers have optimized the perfor-
mance of microorganisms in bioprocessing. Research institutions and collaborative
efforts have been instrumental in advancing nanobiotechnological tools and knowl-
edge, supported by government agencies and organizations. Commercial enterprises
play a crucial role in scaling up and implementing nanobiotechnological solutions
in industrial applications. The combined efforts of these key players hold the poten-
tial to unlock the full potential of biomass resources, contributing to a greener, more
sustainable, and economically viable future. Embracing nanobiotechnological-
based biomass valorization is a crucial step toward achieving a more resource-effi-
cient and environmentally friendly society.
Nanotechnological Advancements for Enhancing Lignocellulosic Biomass Valorization 107

7 Nanotechnology in Lignocellulose Biomass Pretreatment

Pretreatment is a crucial step in the valorization of lignocellulosic biomass, as it


enhances the accessibility of cellulose and hemicellulose to enzymatic hydrolysis.
Conventional pretreatment methods often involve the use of harsh chemicals and
high temperatures, which can lead to significant energy consumption and environ-
mental concerns. Nanotechnology offers innovative solutions to this problem.
Nanoparticles, such as metal oxides (e.g., titanium dioxide, iron oxide), metal
nanoparticles (e.g., gold, silver), carbon-based nanomaterials (e.g., carbon nano-
tubes, graphene), magnetic nanoparticles, and nanocellulose have shown great
promise in breaking down the complex and recalcitrant structure of lignocellulosic
biomass (Khan et al., 2019; Patel et al., 2019; Sukhanova et al., 2018) (Fig. 3).
These nanomaterials increase the surface area and accessibility of biomass, enabling
more efficient enzymatic hydrolysis and subsequent conversion processes.
Nanoparticles modify the biomass surface, break down lignin, and disrupt the ligno-
cellulosic structure, thereby increasing the accessibility of enzymes to cellulose and
hemicellulose (Guisbiers et al., 2012). Also, nanomaterials facilitate the removal of
inhibitory compounds and reduce energy consumption during pretreatment, making
the entire process more environmentally friendly and economically viable.
Additionally, nanocarriers can be used to deliver pretreatment agents more effi-
ciently, reducing their dosage and overall environmental impact.
Pretreatment is a crucial step in the lignocellulosic biomass conversion process.
It aims to disrupt the lignin-carbohydrate complex and increase the accessibility of
cellulose and hemicellulose to enzymatic hydrolysis. Nanotechnology has intro-
duced novel and efficient pretreatment techniques that enhance the overall effi-
ciency of the valorization process. Ionic liquids (ILs) are designer solvents with

Metal
Nanoparticle

Nanomaterials
in
Lignocellulose
Biomass
Valorization

Carbon based Metal Oxide


nanomaterials Nanoparticle

Fig. 3 Nanomaterials in lignocellulose biomass valorization


108 V. Ghosh

unique physicochemical properties that can efficiently dissolve lignocellulosic bio-


mass. Nanoparticle-enhanced ILs have demonstrated superior lignin removal capa-
bilities, thereby reducing the recalcitrance of biomass and facilitating enzymatic
hydrolysis. Metal or metal oxide nanoparticles have been utilized as catalysts in
pretreatment processes. These nanoparticles can selectively break down lignin,
thereby reducing its inhibitory effects on enzyme accessibility to cellulose and
hemicellulose. Nanocatalysts have shown significant potential in improving the effi-
ciency of pretreatment methods. Nanofibrillated cellulose (NFC) derived from
cellulose-­rich biomass has exhibited excellent potential as an effective pretreatment
agent. It can disrupt the lignin-carbohydrate complex and create a more accessible
surface area for enzyme action, leading to higher sugar yields during hydrolysis.
The chemistry of biomass may be improved at a molecular level by using nano-
materials for pretreatment (Rai et al., 2019). Reusability may be improved by utiliz-
ing magnetic nanoparticles, which are easier to remove from the reaction medium
and hence more cost-effective as a whole. As with chemical pretreatment, the
hydrolytic action of nanoparticles or their microemulsions has been shown for LB
processing (Rai et al., 2017). The increasing needs for energy around the world are
satisfied by the utilization of fossil fuels and the application of appropriate technol-
ogy, such as nanotechnology, which provides a viable solution to the problems asso-
ciated with conventional pretreatment procedures.

8 Nano-Enzymatic Systems for Enhanced Hydrolysis

Enzymatic hydrolysis is a key process in lignocellulosic biomass valorization,


where cellulose and hemicellulose are broken down into fermentable sugars.
However, the efficiency of enzymatic hydrolysis is often hindered by nonproductive
adsorption of enzymes onto lignin and limited accessibility to the substrate.
Nanotechnology offers innovative approaches to improve enzymatic hydrolysis
efficiency. Nanobiotechnology has enabled the development of enzyme systems
with enhanced stability, reusability, and catalytic activity. Nanozymes are nanoma-
terials with intrinsic enzyme-like activity which have emerged as a novel class of
catalysts for biomass conversion. These nanozymes can mimic the activity of natu-
ral enzymes, providing advantages such as reduced enzyme loading and improved
resistance to harsh process conditions. With the integration of nanobiotechnological
approaches, enzyme-based biomass conversion processes are becoming more effi-
cient and cost-effective. Nanocarriers, such as nanoparticles or nanogels, have been
employed to immobilize enzymes. This immobilization not only enhances enzyme
stability and recyclability but also improves the efficiency of the enzymatic hydro-
lysis process by increasing the local enzyme concentration on the biomass surface.
Nanocarriers are used for enzyme delivery, protecting enzymes from deactivation
and enhancing their catalytic efficiency (Kaur, 2021). Nanotechnology has enabled
the manipulation of enzyme structures at the nanoscale, leading to engineered
enzymes with improved thermal stability, substrate specificity, and catalytic
Nanotechnological Advancements for Enhancing Lignocellulosic Biomass Valorization 109

activity. These engineered enzymes have shown promising results in breaking down
lignocellulosic biomass more efficiently. Nanocatalysts can be immobilized onto
the surface of cellulases, improving their stability and activity. Functionalized
nanoparticles can also be employed to modify the lignin structure, reducing its
inhibitory effects on enzymatic hydrolysis. These nano-enzymatic systems have
shown promising results in accelerating the hydrolysis process and increasing the
yield of fermentable sugars (Srivastava et al., 2014, 2015). Nanocellulose is derived
from cellulose nanofibrils and has unique properties, such as a large surface area
and high mechanical strength. Nanocellulose has been used as a reinforcing agent
in biocomposite materials for biofuel production. It enhances the mechanical prop-
erties of biofuels, making them more suitable for applications in the automotive and
aviation industries. Magnetic nanoparticles, such as iron oxide (Fe3O4) nanoparti-
cles, have been used to immobilize enzymes for enzymatic hydrolysis of biomass.
By attaching enzymes to the surface of magnetic nanoparticles, they can be easily
recovered and reused in subsequent hydrolysis reactions. This approach reduces
enzyme costs and enhances the efficiency of biofuel production. Zeolites are nano-
porous materials with high surface areas and ion-exchange capabilities. Nanozeolites
have been used as catalysts in the cracking of lignocellulosic biomass to produce
bio-oils. The porous structure of nanozeolites facilitates the conversion of complex
biomass molecules into smaller and more valuable bio-oil fractions. Nanogels are
cross-linked polymeric nanoparticles with tunable properties. They have been used
to encapsulate enzymes and protect them from harsh conditions during enzymatic
hydrolysis of biomass. Nanogels enable controlled enzyme release, improving the
efficiency of the hydrolysis process. Nanofibers, such as carbon nanofibers and sil-
ica nanofibers, have been used as electrode materials in biofuel cells. These nanofi-
bers have high electrical conductivity, which enhances the electron transfer in
biofuel cells and increases their overall efficiency. Nanosensors have been devel-
oped to monitor the fermentation process in biofuel production. These sensors can
detect changes in pH, temperature, and gas concentrations, providing real-time
feedback and control over the fermentation process, leading to improved biofuel
yields. Advanced imaging techniques at the nanoscale level have provided insights
into the enzymatic hydrolysis process. Understanding the interactions between
enzymes and the biomass surface at the nanoscale has led to the design of more
effective enzymatic cocktails and process conditions.
Perfluoroalkylsufonic acid and alkylsulfonic acid functionalized magnetic NP
treatment of wheat straw resulting in a 46% increase in sugar production when com-
pared to the control treatment (35%). Perfluoroalkylsufonic acid and alkylsulfonic
acid might have stabilized and dispersed hemicellulose hydrolysis with an acidity
level comparable to sulfuric acid solutions (Wang et al., 2012). Using magnetic
nanoparticles has a number of advantages, according to the authors, the most nota-
ble of which is their remarkable ability to pretreat lignocellulosic biomass with
small amounts of material while remaining recyclable for future applications. The
combined impact of alkaline pretreatment with magnetite NPs on rice straw for
biogas generation was also identified (Khalid et al., 2019).
110 V. Ghosh

9 Nanomaterials in Biofuel Production

Nanotechnology has also revolutionized the field of biofuel production from ligno-
cellulosic biomass. Biofuels, such as bioethanol and biodiesel, are essential alterna-
tives to fossil fuels, mitigating greenhouse gas emissions and reducing dependency
on nonrenewable resources. Nanomaterials play a vital role in optimizing the vari-
ous stages of biofuel production (Table 1).
For instance, in the fermentation process for bioethanol production, nanocata-
lysts can improve the efficiency of enzymatic or microbial conversion of sugars into
ethanol. Nanomaterials can also be employed as efficient catalysts in biodiesel pro-
duction, converting triglycerides into biodiesel through transesterification reactions.
Additionally, nanosensors and nanoprobes can be utilized for real-time monitoring
and control of biofuel production processes, ensuring optimal conditions and
enhancing overall productivity.

10 Conclusion

Nanotechnological advancements have demonstrated tremendous potential in


enhancing lignocellulosic biomass valorization. Nanotechnology demonstrates
diverse applications of nanoparticles in various stages of biofuel production from
biomass pretreatment to enzymatic hydrolysis and microbial fermentation.
Nanoparticles offer unique properties and capabilities that contribute to the advance-
ment and optimization of biofuel production processes, making them a valuable tool
in the transition to a more sustainable and eco-friendly energy future. Through inno-
vative pretreatment methods and nano-enzymatic systems, nanotechnology facili-
tates the efficient conversion of lignocellulosic biomass into valuable products,
including biofuels and bio-based chemicals. Furthermore, the sustainable nature of
nanomaterials contributes to a more eco-friendly and economically viable approach

Table 1 Application of nanoparticles to improve biofuel production


S. no. Nanoparticle Product Materials References
1. Nickel and Biogas and Weeds Tahir et al. (2020)
cobalt biodiesel
2. NiO Bioethanol Potato peels Sanusi et al. (2021)
3. Fe3O4 Biohydrogen Sugarcane straw Srivastava (2022)
4. Fe3O4 and SiO2 Biohydrogen Sorghum sweet Stover Shanmugam et al. (2020)
5. FeO Biohydrogen Grasses straw Yang and Wang (2018)
6. Nickel Biohydrogen Fermentation Zhang et al. (2021)
7. NiO and CoO Biohydrogen Industrial effluent Mishra et al. (2018)
8. Magnetite Biohydrogen Sugarcane juice Reddy (2017)
9. CaO Biodiesel Sunflower oil Veljković et al. (2009)
10. Ni and Co Biodiesel Chenopodium album Ali et al. (2020)
Nanotechnological Advancements for Enhancing Lignocellulosic Biomass Valorization 111

to biomass valorization. As research and development in nanotechnology continue


to progress, we can expect even more breakthroughs in the field, paving the way for
a more sustainable and greener future. However, it is essential to address any poten-
tial environmental and health concerns associated with the use of nanomaterials to
ensure the responsible and safe deployment of these advancements. Overall, nano-
technology holds the promise of revolutionizing the bio-based industry and acceler-
ating the transition toward a more sustainable and renewable economy.

References

Adewuyi, A. (2022). Underutilized lignocellulosic waste as sources of feedstock for biofuel


production in developing countries. Frontiers in Energy Research, 10, 1–21. https://doi.
org/10.3389/fenrg.2022.741570
Ali, S., Shafique, O., Mahmood, S., Mahmood, T., Khan, B., & Ahmad, I. (2020). Biofuels produc-
tion from weed biomass using nanocatalyst technology. Biomass and Bioenergy, 139, 105595.
https://doi.org/10.1016/j.biombioe.2020.105595
Amthor, J. (2003). Efficiency of lignin biosynthesis: A quantitative analysis. Annals of Botany,
91(6), 673–695. https://doi.org/10.1093/aob/mcg073
Beisl, S., Friedl, A., & Miltner, A. (2017). Lignin from micro- to nanosize?: Applications.
International Journal of Molecular Sciences, 18(11), 2367. https://doi.org/10.3390/
ijms18112367
Bhatia, S. (2020). Recent developments in pretreatment technologies on lignocellulosic bio-
mass: Effect of key parameters, technological improvements, and challenges. Bioresource
Technology, 300, 122724. https://doi.org/10.1016/j.biortech.2019.122724
Chen, H. (2017). A review on the pretreatment of lignocellulose for high-value chemicals. Fuel
Processing Technology, 160, 196–206. https://doi.org/10.1016/j.fuproc.2016.12.007
Costa, S., Aguiar, A., Luz, S., Pessoa, A., & Costa, S. (2015). Sugarcane straw and its cellulose
fraction as raw materials for obtainment of textile fibers and other bioproducts.
Dahadha, S., Amin, Z., Lakeh, A., & Elbeshbishy, E. (2017). Evaluation of different pretreatment
processes of lignocellulosic biomass for enhanced biomethane production. Energy & Fuels,
31(10), 10335–10347. https://doi.org/10.1021/acs.energyfuels.7b02045
Elumalai, S., Agarwal, B., Runge, T., & Sangwan, R. (2018). Biofuel and biorefinery technologies.
Garlapati, V. (2020). Circular economy aspects of lignin: Towards a lignocellulose biorefin-
ery. Renewable and Sustainable Energy Reviews, 130, 109977. https://doi.org/10.1016/j.
rser.2020.109977
Guisbiers, G., Mejía-Rosales, S., & Deepak, F. (2012). Nanomaterial properties: Size and shape
dependencies. Journal of Nanomaterials. https://doi.org/10.1155/2012/180976
Jönsson, L., Alriksson, B., & Nilvebrant, N. (2013). Bioconversion of lignocellulose?: Inhibitors
and detoxification. Biotechnology for Biofuels and Bioproducts, 6, 1–10. https://doi.org/10.118
6/1754-­6834-­6-­16
Kaparaju, P., Serrano, M., Thomsen, A., Kongjan, P., & Angelidaki, I. (2008). Bioethanol, bio-
hydrogen and biogas production from wheat straw in a biorefinery concept. Bioresource
Technology, 100(9), 2562–2568. https://doi.org/10.1016/j.biortech.2008.11.011
Kaur, P. (2021). Nanomaterial conjugated lignocellulosic waste: Cost-effective production
of sustainable bioenergy using enzymes, 3. Biotech, 11(11), 480. https://doi.org/10.1007/
s13205-­021-­03002-­4
Khalid, M., Waqas, Z., & Nawaz, I. (2019). Synergistic effect of alkaline pretreatment and mag-
netite nanoparticle application on biogas production from rice straw. Bioresource Technology,
275, 288–296. https://doi.org/10.1016/j.biortech.2018.12.051
112 V. Ghosh

Khan, I., Saeed, K., & Khan, I. (2019). Nanoparticles: Properties, applications and toxicities.
Arabian Journal of Chemistry, 12(7), 908–931. https://doi.org/10.1016/j.arabjc.2017.05.011
Lo, C., Chang, Y., Chen, Y., Liu, Y., Wu, H., & Sun, Y. (2021). Lignin recovery from rice straw bio-
refinery solid waste by soda process with ethylene glycol as co-solvent. Journal of the Taiwan
Institute of Chemical Engineers, 126, 50–56. https://doi.org/10.1016/j.jtice.2021.07.030
Mishra, P., Thakur, S., Mahapatra, D., Wahid, Z., Liu, H., & Singh, L. (2018). Impacts of nano-­
metal oxides on hydrogen production in anaerobic digestion of palm oil mill effluent – A
novel approach. International Journal of Hydrogen Energy, 43(5), 2666–2676. https://doi.
org/10.1016/j.ijhydene.2017.12.108
Patel, K., Singh, R., & Kim, H. (2019). Carbon-based nanomaterials as an emerging platform for
theranostics. Materials Horizons, 6(3), 434–469. https://doi.org/10.1039/C8MH00966J
Rai, A., Al Makishah, N., Wen, Z., Gupta, G., Pandit, S., & Prasad, R. (2022). Recent develop-
ments in lignocellulosic biofuels, a renewable source of bioenergy. Fermentation, 8(4), 161.
https://doi.org/10.3390/fermentation8040161
Rai, M., Ingle, A., Gaikwad, S., Dussán, K., & da Silva, S. (2017). Nanotechnology for bioenergy
and biofuel production.
Rai, M., Ingle, A., Pandit, R., Paralikar, P., Biswas, J., & da Silva, S. (2019). Emerging role of
nanobiocatalysts in hydrolysis of lignocellulosic biomass leading to sustainable bioethanol
production. Catalysis Reviews: Science and Engineering, 61(1), 1–26. https://doi.org/10.108
0/01614940.2018.1479503
Reddy, K. (2017). Biohydrogen production from sugarcane bagasse hydrolysate: Effects of pH,
S/X, Fe2+, and magnetite nanoparticles. Environmental Science and Pollution Research, 24,
8790–8804. https://doi.org/10.1007/s11356-­017-­8560-­1
Rezania, S. (2020). Different pretreatment technologies of lignocellulosic biomass for bioethanol
production: An overview. Energies, 199, 117457. https://doi.org/10.1016/j.energy.2020.117457
Saini, J., Saini, R., & Tewari, L. (2015). Lignocellulosic agriculture wastes as biomass feedstocks
for second-generation bioethanol production?: Concepts and recent developments, 3. Biotech,
5, 337–353. https://doi.org/10.1007/s13205-­014-­0246-­5
Sankaran, R. (2021). The expansion of lignocellulose biomass conversion into bioenergy
via nanobiotechnology. Frontiers in Nanotechnology, 3, 1–10. https://doi.org/10.3389/
fnano.2021.793528
Sanusi, I., Suinyuy, T., & Kana, G. (2021). Impact of nanoparticle inclusion on bioethanol pro-
duction process kinetic and inhibitor profile. Biotechnology Reports, 29, e00585. https://doi.
org/10.1016/j.btre.2021.e00585
Scarlat, N., Dallemand, J., Monforti-ferrario, F., & Nita, V. (2015). The role of biomass and bio-
energy in a future bioeconomy?: Policies and facts. Environment and Development, 15, 3–34.
https://doi.org/10.1016/j.envdev.2015.03.006
Shanmugam, S., Krishnaswamy, S., Chandrababu, R., Veerabagu, U., Pugazhendhi, A., &
Mathimani, T. (2020). Optimal immobilization of Trichoderma asperellum laccase on polymer
coated Fe3O4@SiO2 nanoparticles for enhanced biohydrogen production from delignified lig-
nocellulosic biomass. Fuel, 273, 117777. https://doi.org/10.1016/j.fuel.2020.117777
Srivastava, N. (2022). Biohydrogen production via integrated sequential fermentation using magne-
tite nanoparticles treated crude enzyme to hydrolyze sugarcane bagasse. International Journal
of Hydrogen Energy, 47(72), 30861–30871. https://doi.org/10.1016/j.ijhydene.2021.08.198
Srivastava, N., Rawat, R., Sharma, R., Oberoi, H., Srivastava, M., & Singh, J. (2014). Effect of
nickel cobaltite nanoparticles on production and thermostability of cellulases from newly iso-
lated thermotolerant Aspergillus fumigatus NS (Class: Eurotiomycetes). Applied Biochemistry
and Biotechnology, 174, 1092–1103. https://doi.org/10.1007/s12010-­014-­0940-­0
Srivastava, N., Singh, J., Ramteke, P., Mishra, P., & Srivastava, M. (2015). Improved production of
reducing sugars from rice straw using crude cellulase activated with Fe3O4/Alginate nanocom-
posite. Bioresource Technology, 183, 262–266. https://doi.org/10.1016/j.biortech.2015.02.059
Sukhanova, A., Bozrova, S., Sokolov, P., Berestovoy, M., Karaulov, A., & Nabiev, I. (2018).
Dependence of nanoparticle toxicity on their physical and chemical properties. Nanoscale
Research Letters, 13, 1–21. https://doi.org/10.1186/s11671-­018-­2457-­x
Nanotechnological Advancements for Enhancing Lignocellulosic Biomass Valorization 113

Tahir, N., Tahir, M., Alam, M., Yi, W., & Zhang, Q. (2020). Exploring the prospective of weeds
(Cannabis sativa L., Parthenium hysterophorus L.) for biofuel production through nano-
catalytic (Co, Ni) gasification. Biotechnology for Biofuels, 13, 148. https://doi.org/10.1186/
s13068-­020-­01785-­x
Valdivia, M., Galan, J., Laffarga, J., & Ramos, J. (2016). Biofuels 2020: Biorefineries based on
lignocellulosic materials. Applied Microbiology International, 9(5), 585–594. https://doi.
org/10.1111/1751-­7915.12387
Veljković, V., Stamenković, O., Todorović, Z., Lazić, M., & Skala, D. (2009). Kinetics of sun-
flower oil methanolysis catalyzed by calcium oxide. Fuel, 88(9), 1554–1562. https://doi.
org/10.1016/j.fuel.2009.02.013
Wang, D., Ikenberry, M., Peña, L., & Hohn, K. (2012). Acid-functionalized nanoparticles for
pretreatment of wheat straw. Journal of Biomaterials and Nanobiotechnology, 3(2), 342–352.
https://doi.org/10.4236/jbnb.2012.33032
Yang, G., & Wang, J. (2018). Improving mechanisms of biohydrogen production from grass
using zero-valent iron nanoparticles. Bioresource Technology, 266, 413–420. https://doi.
org/10.1016/j.biortech.2018.07.004
Zhang, J., Zhao, W., Yang, J., Li, Z., Zhang, J., & Zang, L. (2021). Comparison of mesophilic
and thermophilic dark fermentation with nickel ferrite nanoparticles supplementation for
biohydrogen production. Bioresource Technology, 329, 124853. https://doi.org/10.1016/j.
biortech.2021.124853
A State of the Art of Biofuel Production
Using Biomass Wastes: Future Perspectives

Thi An Hang Nguyen, Thi Viet Ha Tran, and Minh Viet Nguyen

1 Introduction

There is a global growing interest in recycling biomass wastes into biofuels as


depicted in Fig. 1. This is driven by rising energy demand, fossil fuel depletion,
and climate change (Lee et al., 2023). It also results from an elevated demand
for waste disposal, which otherwise causes environmental burdens (Sonu
et al., 2023).
Various biomass wastes can be used for biofuel production, including agrofor-
estry wastes (known as lignocellulosic biomass) (Beltrán-Ramírez et al., 2019), ani-
mal manures (Jung et al., 2021), industrial wastes (Gil, 2022), and organic fraction
of municipal solid wastes (MSW) (Yu et al., 2023). Each category of biomass waste
has its own properties, influencing subsequent conversion processes (Irmak, 2019).
Depending on a specific biomass waste and desired bioenergy end products (e.g.,
solid, liquid, gaseous biofuels, or heat), different pretreatment methods can be
applied, such as physical, chemical, physicochemical, and biological methods
(Nadir et al., 2019).
To convert biomass wastes into biofuels, numerous technologies have been
examined that can be classified as physicochemical, thermochemical, and biochem-
ical methods (Anekwe et al., 2022). While conventional technologies present sig-
nificant limitations, emerging technologies offer several advantages (e.g., high

T. A. H. Nguyen (*) ∙ T. V. H. Tran


Faculty of Advanced Technologies and Engineering, VNU Vietnam Japan University,
Hanoi, Vietnam
M. V. Nguyen
VNU Key Laboratory of Advanced Material for Green Growth, VNU University of Science,
Hanoi, Vietnam

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 115
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_6
116 T. A. H. Nguyen et al.

Fig. 1 Number of published articles on biomass waste, biofuel, and both keywords in the period
of 2012–2022 according to the ScienceDirect database

energy efficacy, less inhibitors formation, marginal environmental impacts, low car-
bon footprint) (Kumar et al., 2022). To enhance the techno-economic viability of
existing WtE technologies, applying of co-feedstocks (Fang et al., 2017), advanced
pretreatment methods (Al Ramahi et al., 2021), integrated conversion technologies,
and tuning of process parameters (Yu et al., 2023) are proven to be effective
solutions.
This chapter provides an overview of the diversity of biofuels, major types of
biomass wastes, numerous WtE conversion technologies, process enhancement
strategies, and future perspectives.

2 Diversity of Biofuels

2.1 Biofuel Generations

There are four biofuel generations, depending on feedstock sources. The first-­
generation biofuels are produced from food crops (e.g., corn, sugarcane, soybean)
using traditional techniques. Bioalcohols, biodiesels, biogas, bioethers, biosyngas,
and vegetable oil are examples of the first-generation biofuels. Contrarily, the
second-­generation biofuels derive from nonfood-based materials (e.g., agroforestry
wastes), including biohydrogen and bioethanol. The third-generation biofuels are
algae-based biofuels, namely, biohydrogen and biomethane (Anekwe et al., 2022).
The fourth-generation biofuels originate from industrial waste or metabolic prod-
ucts of algae (Sonu et al., 2023).
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 117

2.2 Typical Biofuels

Biofuels originate from biomass and can exist in solid, liquid, or gas phases
(Gil, 2022).
Solid biofuels are pellets produced from agroforestry residues and wood indus-
tries by a several-step procedure, namely, drying, milling, pressing, packaging, and
storing. To improve the energy density of pellets, several techniques have been used,
such as briquetting, torrefaction (Sahoo et al., 2019), steam explosion, hydrothermal
carbonization (HTC) (Kang et al., 2019), and biological treatment (Beltrán-Ramírez
et al., 2019).
Liquid biofuels comprise biodiesel, bio-oil, bioethanol, and butanol (Beltrán-­
Ramírez et al., 2019). Biodiesel is made from animal fats, vegetable oils, and waste
cooking oil. It can substitute for diesel because of its nontoxicity, biodegradability,
no sulfur, and benzene containing. Biodiesel is produced via a three-step process
including pretreatment, transesterification, and separation. Pretreatments are divided
into acidic, basic, thermal, enzymatic, or combined treatments. Bio-oil is a mixture
of organic compounds (e.g., acids, alcohols, aldehydes, esters, ketones, and phe-
nols) attained via pyrolysis or liquefaction. Fast pyrolysis offers low cost, whereas
liquefaction produces low yield at excessive cost. Pyrolysis of agro-industrial
wastes (e.g., mustard, palm kernel, cottonseed, and neem oil cakes) helps to valorize
them and prevent environmental pollution. For the application of bio-oils as biofu-
els, pretreatments are required, including cracking and hydrotreating to remove O2.
Since bio-oils emit less CO2 and NOx than diesel, it is considered a green fuel.
Bioethanol is the most widely used alcoholic biofuel as it has renewable agricultural
feedstocks and is environmentally benign. It is produced via a four-step process,
including pretreatment, saccharification, fermentation, and distillation. Butanol
presents advantages, namely, high heat of combustion, less volatility, and blending
ability with gasoline without engine modification requirement. It is produced via
anaerobic fermentation using Clostridia genus (Anekwe et al., 2022).
Gaseous biofuels consist of biogas and biohydrogen. Biogas is generated from
agricultural, animal, green, food wastes, etc., via anaerobic digestion (AD) process.
Biogas (mainly CH4 and CO2) is used for heat or electricity generation. Biogas pro-
duction involves four steps, including hydrolysis, acidification, acetate formation,
and methane (CH4) generation. To intensify biogas formation, pretreatments are
needed, including acid hydrolysis, steam explosion, alkaline hydrolysis, and liquid
hot water (Beltrán-Ramírez et al., 2019). Biohydrogen is formed via gasification or
microbial fermentation of biomass. Gasification presents significant drawbacks,
such as intensive energy consumption and high emissions (e.g., C, S, NOx).
Contrarily, microbial fermentation is more eco-friendly and can be divided into dark
fermentation, photo-fermentation, and sequential dark-photo-fermentation.
Biohydrogen is mainly produced at laboratory scale. Biohydrogen production can
be enhanced by employing pretreatments of feedstocks. Factors governing biohy-
drogen formation are feedstock categories, pretreatment methods, and types of
microorganisms (Salakkam et al., 2019).
118 T. A. H. Nguyen et al.

3 Biomass Wastes as Feedstocks for Biofuel Production

Biomass is the plant or animal derived from organic matters (e.g., sugars, starches,
lignocelluloses). It is a potential source of renewable bioenergy due to its plentiful
supply, low cost, and low greenhouse gas (GHG) emission (Chandraratne & Daful,
2022). Converting biomass wastes into biofuels results in many benefits, such as i)
minimizing the requirement of biomass waste disposal, (ii) reducing environmental
risks (e.g., water contamination, odor pollution, GHG emission, pests, and insect
breeding), and (iii) generating alternative energy to fossil fuels (Chandraratne &
Daful, 2022). Biomass wastes include lignocellulosic, livestock and solid municipal
and industrial wastes.

3.1 Lignocellulosic Wastes

Lignocellulosic waste is a major feedstock for biofuel production, comprising com-


plex biopolymers, e.g., cellulose, hemicellulose, and lignin. The annual lignocellu-
losic waste worldwide was 181.5 billion tons (Lee et al., 2023). It can be categorized
as (i) field/crop residues, e.g., rice straw (Asadi & Zilouei, 2017), corn stalks (Kang
et al., 2019), and defective coffee beans (Santana et al., 2020); (ii) forest residues,
e.g., sawdust (Czekała et al., 2018) and woodchips (Sahoo et al., 2019); and (iii)
industrial wastes, e.g., edible oil waste (Amenaghawon et al., 2021), fruits and veg-
etables processing waste (Edwiges et al., 2018), apple pulp waste (Gökçek et al.,
2023), brewery processing waste (Panjičko et al., 2017), potato waste (Sekoai et al.,
2019), and rice bran (Tandon et al., 2018). Various biofuels can be produced from
lignocellulosic wastes including (i) liquid biofuels (e.g., bioethanol, biodiesel,
ether) and (ii) gaseous biofuels (e.g., biogas, biohydrogen) (Sonu et al., 2023).
Amenaghawon et al. (2021) produced biodiesel by transesterification of edible oil
wastes at 1200 °C with a biodiesel yield of 59.89–98.54%. To date, the use of ligno-
cellulosic wastes for biofuel production still faces difficulties due to their properties
(e.g., complex structure, recalcitrance, dissolution difficulty, low density, high
moisture) (Irmak, 2019). Besides, there are other challenges, such as availability,
storage, stable supply, affordable cost, and uniformity (Clauser et al., 2021).

3.2 Livestock Manures

Livestock manures are characterized by the following chemical constituents: cellu-


lose (7.7–42.4%), hemicellulose (9.2–31.4%), lignin (3–14.5%), protein
(16–48.4%), lipid (2.6–22%), ash (8.7–15.7%), and other inorganic matters (Jung
et al., 2021). Thanks to their diverse chemical constituents (e.g., protein, lipid, car-
bohydrates) and massive production, livestock manures can be used for the
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 119

production of various biofuels (e.g., heat, biogas, bioethanol, biodiesel). Though


livestock manures can be directly used for heat generation via burning, they demon-
strate a substantially lower higher heating value (HHV) (12–19 MJ/kg) than bio-
ethanol (29.7 MJ/kg) or biodiesel (38–41 MJ/kg). Livestock manures are also
employed for bioethanol production. The lower degree of polymerization and
smaller particle sizes of livestock manures compared to lignocellulosic wastes favor
sugar formation (Jung et al., 2021). Moreover, as livestock manures are rich in lipids
(11–14%), they are applied in biodiesel production via two steps, namely, lipid
extraction and transesterification (Jung et al., 2021). It was reported by Kim et al.
(2020) that by transesterification of swine manure with an acidic catalyst (H2SO4) at
60 °C for 24 h, the obtained biodiesel yield was 14.2% based on lipid content. This
yield is lower than that of chicken, goat manure (35.7%), and cow manure (54.1%)
(Gomaa & Abed, 2017). Besides, livestock manures are utilized for biogas genera-
tion via AD. The biogas and biomethane yields from livestock manures were in the
ranges of 0.16–0.42 and 0.07–0.36 m3/kg VS, respectively (Jung et al., 2021).
According to the U.S. EIA (2019), livestock manures are the second largest feed-
stocks for biogas production in Europe (6.1 MTOE).
Livestock manures also exhibit limitations for biofuel production. First, high
moisture content (up to 80%) leads to intensive energy consumption for pre-drying,
thus increasing the cost of ethanol. It also affects transesterification of lipids to pro-
duce biodiesel. It was recommended to integrate AD with fermentation to enhance
both bioethanol and biogas production. Second, high nitrogen content (>2%)
adversely influences acid hydrolysis. Specifically, the raw dairy manure (2.6% N)
resulted in a 10–20% lower yield of pentose than the pretreated dairy manure (1.3%
N) in acid hydrolysis. Finally, the C/N ratios of livestock manures (<15) are lower
than the optimal C/N ratios for biogas production (20–30). This is overcome by co-­
digestion of livestock manures with lignocellulosic wastes (Jung et al., 2021). So
far, less attention has been paid to livestock manures than lignocellulosic materials
as feedstocks for biofuel production.

3.3 Organic Fraction of Municipal Solid Waste (MSW)

MSW is an attractive feedstock for biofuel production. Around 1.9 billion tons of
MSW are generated worldwide annually. MSW composition varies, depending on
locations. The typical composition of MSW is organic waste (46%), wastepaper
(17%), plastics (10%), glass (5%), metal (4%), and miscellaneous (18%) (Nanda &
Berruti, 2021).
MSW can be converted into biofuels using such technologies as (i) incineration,
(ii) thermochemical processes, and (iii) biological processes. Typical biofuels (con-
version technology) are heat (incineration), biochar, bio-oil, and syngas (thermo-
chemical processes), biogas (anaerobic digestion), compost (composting),
biohydrogen (dark and photo-fermentation, bio-photolysis, microbial electrolysis
cells, and microbial electro-hydrogenesis cells). Incineration reduces the weight
120 T. A. H. Nguyen et al.

and volume of MSW by 80–85% and 95–96%, respectively, mitigates odor pollu-
tion, and prevents the development of rodents and flies. However, incomplete incin-
eration induces air pollution. Hydrothermal processes present advantages (e.g.,
short reaction time, negligible toxic residues, high energy recovery). Aurnob et al.
(2022) observed that HTC of MSW resulted in the solid yield of 48.98%, HHV of
15.4 MJ/kg, and energy retention rate of 69.3%. Scarlat et al. (2015) reported that
one ton of wet MSW (60% organic matter and 40% moisture) generated 150 kg of
CH4 via the AD process. However, toxins in MSW can inhibit methanogens, thereby
impeding methanogenesis (Nanda & Berruti, 2021).
To maximize biofuel production from MSW, it is crucial to segregate MSW into
different components for disposal with proper technology. Nanda & Berruti (2021)
stated that organic fraction in MSW (e.g., kitchen, food, and yard wastes) is suitable
for biological processes (e.g., AD, composting) and hydrothermal processes (e.g.,
liquefaction, HTC). In contrast, recalcitrant organic components (e.g., paper waste,
packaging boxes, cardboards) and nonbiodegradable organics (e.g., plastics, rubber,
tires) are compatible with thermochemical processes (e.g., pyrolysis, gasification).

3.4 Industrial Biomass Wastes

Various organic wastes from industries have been used for biofuel production. They
can be divided into (i) non-lignocellulosic materials (e.g., seafood waste), (ii) ligno-
cellulosic materials (e.g., sugar baggage, apple pulp waste), and (iii) ash-rich mate-
rials (e.g., sewage sludge). Recycling industrial biomass wastes into biofuels may
encounter challenges, such as high moisture (up to 95%), low energy density, and
toxin contamination. Abundant water content in industrial biowastes can be attrib-
uted to the formation of extracellular polymeric substances (EPSs) on their surface
to retain moisture (Zhuang et al., 2022). Among conversion technologies, HTC
presents merits, such as improved dewaterability efficiency (Wang et al., 2017),
enhanced fuel properties (Liu et al., 2013), reduced NOx, SOx, and HCl emissions
(Indrawan et al., 2012), and increased stabilization of heavy metals in industrial
biowastes. Additionally, HTC can be used as pretreatment for other processes to
enhance bioenergy recovery. Moon et al. (2015) reported that the HTC of sewage
sludge improved syngas quality and reduced its moisture to facilitate the subsequent
pyrolysis.

4 Technologies for Conversion of Biomass Wastes


into Biofuels

Numerous extractions, thermochemical, biochemical conversion, and hybrid pro-


cesses have been developed for the conversion of biomass wastes into biofuels as
shown in Fig. 2 on both small scale and large scale (Shafizadeh & Danesh, 2022).
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 121

Fig. 2 Conversion technologies for biofuel production from biowastes. (Modified from Anekwe
et al., 2022; Shafizadeh & Danesh, 2022)

4.1 Physicochemical Conversion Processes

Extraction Extraction is applied to extract oil from oil seeds or nuts (e.g., sesame
seed, sunflower seed, rapeseed, hazelnut, almond nut). This is done with traditional
or advanced methods. Traditional methods comprise mechanical methods (e.g., cold
pressing, hot pressing) and chemical methods (e.g., solvent extraction). Normal
extraction has three steps. The first pretreatment step is to improve extraction effi-
ciency by destroying/softening the cellular structure and reducing the moisture con-
tent in the biowaste. The second pressing step is to extract oil from the pretreated
biowaste to form a press cake. The final chemical extraction step is to extract the
remaining oil (20–30%) from the press cake. The advanced techniques for oil
extraction from solid-liquid samples are microwave-assisted extraction, supercriti-
122 T. A. H. Nguyen et al.

cal fluid extraction, and ultrasound-assisted extraction. So far, extraction is used at


laboratory scale (low and medium levels) (Shafizadeh & Danesh, 2022).

Transesterification Transesterification is a process of biodiesel production, which


can be promoted by both homogenous and heterogenous catalysts. Homogenous
catalysts include alkaline catalysts (e.g., NaOH, KOH) and acid catalysts (e.g.,
H2SO4, HCl). Though alkaline catalysts have significant advantages (e.g., applica-
ble at low temperature and atmospheric pressure, high conversion yield), they are
not suitable for biowastes with high free fatty acids content. This challenge is over-
come by using acid catalysts provided that longer time and higher temperature are
supplied. Despite excellent conversion efficiency, the applicability of homogenous
catalysts is restricted by their weight loss. Therefore, heterogenous catalysts can be
utilized alternatively to enable their recovery and reuse (Anekwe et al., 2022).

4.2 Thermochemical Processes

Thermochemical conversion processes are the degradation of biopolymers in bio-


wastes into biofuels and other chemicals via heat exposure and chemical reactions.
These processes can be classified as pyrolysis, hydrothermal processing, combus-
tion, and gasification. Depending on process conditions, main products may include
solid (biochar, hydrochar), liquid (bio-oil, bio-crude oil), and gaseous (tar, syngas)
forms (Shafizadeh & Danesh, 2022). Thermochemical conversion processes dem-
onstrated several merits as compared to biochemical processes, such as being appli-
cable to various feedstocks and being conducted in a short residence time
(Chandraratne & Daful, 2022).
Pyrolysis It is an exothermic thermal process that depolymerizes biowastes in an
inert atmosphere (in the presence of N2 or absence of O2) and high temperatures
(280–1000 °C) to produce biochar, bio-oil (as the main products), and biosyngas (as
the by-product) (Garba, 2020). A three-step mechanism exists including dehydroge-
nation, depolymerization, and fragmentation (Chandraratne & Daful, 2022).
Pyrolysis performance is influenced by feedstock nature and pyrolysis conditions
(e.g., temperature, pressure, heating rate, residence time, environment, catalyst).
The feedstock nature (e.g., cellulose, hemicellulose, lignin) affects pyrolytic behav-
iors, causing the difference in product composition. The low temperature enables
the formation of solid fraction. On the contrary, elevated temperatures and short
residence times enhance the production of condensable fraction, whereas elevated
temperatures and long residence times favor the formation of non-condensable
gases (Garba, 2020). Based on operational conditions and desired products, pyroly-
sis can be classified as fast, intermediate, and slow pyrolysis. Slow pyrolysis is
conducted at temperatures of 300–700 °C, heating rates of 0.1–1 °C/s, and resi-
dence times of 10–100 min, resulting in more biochar (Nanda & Berruti, 2021). A
slow and mild pyrolysis, which is operated at temperatures of 450–550 °C, heating
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 123

rates of <1 °C/s, for less than 2 h to remove moisture and organic volatiles from
feedstocks to attain torrefied materials, is called torrefaction (Chandraratne & Daful,
2022). Intermediate pyrolysis is conducted at temperatures of 500–600 °C, heating
rates of 2–10 °C/s, and residence times of 10–20 min, resulting in a low liquid yield
(up to 55%). Fast pyrolysis is operated at temperatures of 400–500 °C, heating rates
of 10–200 °C/s, and residence times of 0.03–1.5 s, producing more bio-oil (up to
75%) (Nanda & Berruti, 2021). The HHV of bio-oils varies in the range of 15–20 MJ/
kg, being 40–50% that of conventional petroleum oils (42–45 MJ/kg) (Garba,
2020). Flash pyrolysis is conducted at temperatures of 800–1000 °C, heating rates
of 1000 °C/s, and residence times of 0.5 s (Nanda & Berruti, 2021). Pyrolysis is
available at laboratory and commercial scales with semicontinuous or continuous
mode. The advanced pyrolysis technologies include plasma, vacuum, microwave-­
assisted, and solar-assisted pyrolysis, which exhibit merits over traditional electrical-­
heating-­assisted pyrolysis. Regarding the technical readiness level, pyrolysis is
available at pilot and industrial scales (Shafizadeh & Danesh, 2022). Lee et al.
(2021) produced biochar by slow pyrolysis of spent coffee ground wastes at 500 °C
and heating rate of 12 °C/min for 30 min. The resulting biochar had a HHV value of
31.41 MJ/kg.

Hydrothermal Processing Hydrothermal processing can save energy since it does


not require pre-drying of wet biowastes. Therefore, it is especially compatible with
wet biowastes. It can be divided into HTC, hydrothermal liquefaction (HTL), and
hydrothermal gasification (HTG), which aim at producing the largest solid (hydro-
char), liquid (bio-oil/water), and gas fractions, respectively. They differentiate in
terms of such operational parameters as temperature, pressure, and residence time
(Chandraratne & Daful, 2022). HTC is a hydrothermal process conducted with the
presence of water at below 200 °C, autogenous pressure, for 5–240 min, resulting in
hydrochar as the main product. HTL is a thermochemical process performed at
elevated temperatures (200–350 °C) and high pressures (5–20 MPa) in the presence
of solvent (subcritical/supercritical water) to produce bio-crude (the main product),
biochar, water-soluble organic polar fractions, and gases. The resultant bio-crude oil
demonstrates HHV of 36–40 MJ/kg, being equivalent to that of the petroleum-­
derived oil. The quantity and quality of bio-crude oil rely on feedstock composition
and operational parameters (e.g., temperature, reaction time, pressure, catalyst,
water-to-biomass ratio, and reaction medium). HTL occurs via complicated mecha-
nisms including hydrolysis, activation, and depolymerization. In comparison with
other thermochemical processes, HTL offers merits, such as (i) suitable with wet
biowastes and ii) resulting high-quality bio-crude oil thanks to superheated fluid,
solvent, and high pressure. HTL is available at the pilot scale. HTG is defined as a
thermochemical process to convert wet biomass into combustible gases in tempera-
tures of 400–600 °C and pressures of 24–36 MPa with/without catalysts. In HTG,
supercritical water plays the roles of reactant and solvent to break down organic
matters. Consequently, reactive species are dissolved, hence enhancing the yield of
gases, while hindering the biochar formation. So far, hydrothermal processes are
ready at the laboratory scale (high level) and pilot scale (low or medium level)
124 T. A. H. Nguyen et al.

(Shafizadeh & Danesh, 2022). Santana et al. (2020) produced hydrochar via HTC of
defective coffee beans at temperatures of 150, 200, and 250 °C. The energy yield
and HHV of the resultant hydrochar were 46.0–61.0% and 20.20–29.10 MJ/kg,
respectively.

Incineration Incineration is exothermic oxidation of biomass at high temperatures


(800–1200 °C) in the presence of oxygen to form heat as the main product (Nanda
& Berruti, 2021). It consists of four consecutive steps including drying, devolatil-
ization, volatiles combustion, and char combustion. The process products vary,
depending on feedstock properties, particle size, temperature, and combustion air.
The solid products comprise char and ash, whereas the gaseous products include
CO2 and H2O (complete combustion) or carbon monoxide (CO), CH4, non-methane
hydrocarbons (NMHCs), particulate matter (PM), and NOx and SOx (incomplete
combustion). In order to mitigate emissions, combustion is modified by gas recircu-
lation, boiler modification, and re-burning. Combustion is used at industrial scale
(high level) (Shafizadeh & Danesh, 2022).

Gasification It is an endothermic thermochemical process that converts biomass


into combustible gases, known as syngas, with the presence of gasifying agents
(e.g., air, steam, CO2) in elevated temperatures (800–1300 °C) (Garba, 2020).
Gasification has four stages that occur in different temperatures, namely, drying
(100–200 °C), pyrolysis (200–450 °C), reduction (450–650 °C), and gasification
(650–1300 °C). The generated gas has varying composition (e.g., CO, H2, CH4,
CO2, H2O, and N2), depending on biomass properties (e.g., feedstock type, compo-
sition, particle size, moisture, and ash content), and operational conditions (e.g.,
temperature, pressure, residence time, air-to-biomass ratio, catalyst type, and quan-
tity) (Shafizadeh & Danesh, 2022). The high moisture content in biowastes reduces
conversion efficiency via incomplete cracking and oxidation and lowers the produc-
tion rate. It may also affect the syngas composition (e.g., high contents of CO2 and
H2O), requiring additional treatment facilities. Gasification offers advantages
including high flexibility and multiple usages of the generated syngas (Garba,
2020). However, obstacles exist including biowaste collection and storage, biowaste
pretreatment, gas cleaning, process efficiency, and syngas quality. Relating the
­technological maturity, gasification is ready at pilot scale (high level) and industrial
scale (all levels) (Shafizadeh & Danesh, 2022).

4.3 Biochemical Conversion Processes

Anaerobic Digestion
AD is a biochemical process that converts biowastes into biogas in an inadequate O2
medium. AD consists of four consecutive steps, including hydrolysis, acidogenesis,
acetogenesis, and methanogenesis. During the first step, complex and insoluble bio-
polymers are degraded into simple and soluble monomers. Specifically, hydrolysis
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 125

of carbohydrate, protein, and lipid polymers resulted in the formation of sugar,


amino acid, and long-chain fatty acid monomers, respectively. In the second step,
volatile fatty acids (VFAs) are formed from monomers, thanks to acidogenic bacte-
ria. The third step results in the formation of acetic acid and hydrogen (H2) as the
main products. In the last step, CH4 is produced from acetic acid and H2 via aceto-
trophic, hydrogenotrophic, and methylotrophic routes. This leads to a rise in the
system’s pH, reaching the neutral pH range of 6.8–8 (Shafizadeh & Danesh, 2022).
The generated biogas consists of CH4 (60%), CO2 (35%), and other gases, such as
H2, NH3, CO, and H2S (5%). The biogas with the mixing ratio of CH4: CO2: inert of
60: 35: 5 demonstrates the heating value of 22.35 KJ/m3 (Garba, 2020).
The AD performance is influenced by many process parameters, such as tem-
perature, carbon to nitrogen (C: N) ratio, pH, hydraulic retention time (HRT),
organic loading rate (OLR), and stirring. The temperature may affect the stability
and activity of enzymes needed for AD and CH4 formation. The optimal tempera-
ture ranges for diverse groups of anaerobic microorganisms are as follows: 10–30 °C
(psychrophilic), 30–40 °C (mesophilic), and 50–60 °C (thermophilic). Similarly,
the system’s pH can influence the activity of acidogenic and methanogenic bacteria,
which are dominant in acidogenesis and methanogenesis, respectively. The optimal
pH ranges are 5.5–6.5 and 6.5–8.2 for acidogenesis and methanogenesis, respec-
tively. The C/N ratio is another important process parameter. The optimal C/N ratio
is in the range of 20–35. The low C/N ratio may lead to the AD inhibition caused by
high NH4+ level, whereas the high C/N ratio implies the need for nutrient supple-
mentation to sustain the AD process. The HRT can vary, depending on the feedstock
and temperature. While short HRT enhances the AD’s efficacy and reduces the cost,
long HRT improves biogas yields. The high OLR leads to the lack of resources for
microbial growth, whereas low OLR may cause microbial malnutrition. AD is per-
formed with two types of reactors, depending on the solid concentration. Dry and
wet reactors are used for feedstocks with the total solid content above and below
15%, respectively. Dry reactors are divided into horizontal plug flow, vertical plug
flow, and non-flow (batch type). Concerning the technical readiness level, AD is
available at pilot scale (medium or high level) and commercial scale (low or medium
level) (Shafizadeh & Danesh, 2022).
Fermentation
Fermentation is a biochemical process that converts simple sugars into low molecu-
lar alcohols or acids, thanks to microbial enzymes under mild conditions (Garba,
2020). The most commercially used microorganism for fermentation is
Saccharomyces cerevisiae. Fermentation can be implemented with food-based
feedstocks (e.g., wheat, corn, sugarcane, soybean, sunflower) or nonfood-based
feedstocks (e.g., agro-waste, green waste) to produce the first-generation or the
second-generation biofuels, respectively. There are several factors governing fer-
mentation, including temperature, pH, aeration, and nutrient supplementation.
Depending on feedstock composition, pretreatments may be required, which can be
classified as physical, chemical, and biological methods (Shafizadeh & Danesh,
2022). A big challenge to biofuel production from biowastes is the formation of
126 T. A. H. Nguyen et al.

inhibitors along with desired products, thus declining the conversion ratio. This can
be overcome by improving microbial metabolism (Garba, 2020).
Fermentation is implemented in batch, continuous, and fed-batch modes. The
batch mode offers advantages, such as complete and high rate of substrate conver-
sion and high productivity, while having disadvantages including long duration and
lower volumetric production. Contrarily, the continuous mode can use the plant
capacity more efficiently. So as to intensify fermentation, the addition of
Streptococcus sp. or glucoamylase enzyme is useful. Fermentation has the same
technical readiness level as AD, being available at pilot scale (medium or high level)
and industrial scale (low or medium level) (Shafizadeh & Danesh, 2022).

5 Enhancement Strategies for Production of Biofuels


from Biowastes

Though biowaste-derived biofuels can potentially substitute for fossil fuels, their
wide utilization is still hindered by biomass supply chain, energy efficiency, and
product yield. To address these problems, various improvement strategies are used
individually or combinedly (Anekwe et al., 2022).

5.1 Pretreatments of Biomass Wastes

Lignocellulosic biowastes possess unfavorable properties for biofuel production


(e.g., low density, high moisture, structural complexity, recalcitrance, and dissolu-
tion difficulty). Therefore, pretreatment of lignocellulosic biowastes is vital in bio-
fuel production to enhance the conversion efficiency (Irmak, 2019; Kumar et al.,
2022). Pretreatment methods need to satisfy the following requirements: (i) increase
sugar formation, (ii) minimize the formation of inhibitors to subsequent hydrolysis
and fermentation, (iii) reduce the loss of carbohydrates, and (iv) be cost-effective.
Depending on the process’s nature, pretreatment methods of biowastes are catego-
rized as (1) physical, (2) chemical, (3) physicochemical, and (4) biological methods
(Nadir et al., 2019). The objectives, classification, merits, and demerits of biowaste
pretreatment methods are illustrated in Fig. 3.
Another way for classification of pretreatment methods is based on feedstock
categories and desired products, including (i) bioethanol, (ii) biodiesel, and (iii)
other products. For bioalcohol production from sugar- or starch-containing crop
residues, sugar extraction is enhanced via pretreatments including milling, liquefac-
tion, and saccharification. To produce biodiesel from livestock manures and oil
cake, such pretreatments as lipid extraction and separation of impurities are neces-
sary. In order to valorize the organic fraction in lignocellulosic biowastes, food and
paper industry waste as biochar, bio-oil, syngas, and biogas, a series of
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 127

Fig. 3 Objectives, classification, merits, and demerits of pretreatment methods for biofuel produc-
tion from biowastes

pretreatments are performed, namely, separation of lignin, hydrolysis of carbohy-


drates, and detoxification.
Based on the advancement of pretreatment methods, they can be classified as
conventional and emerging technologies. While conventional pretreatment methods
are limited by techno-economic feasibility, carbon footprints, process speed, and
generation of inhibitors, emerging pretreatment processes are advanced, green, and
cost-effective. Emerging pretreatment methods can be grouped into (i) physico-
chemical methods (e.g., microwaving, ultrasound, radiation, steam explosion pre-
treatment (SEP), hydrothermal, pulse-electric field (PEF)), (ii) chemical methods
(e.g., ionic liquids (Ils), deep eutectic solvents (DESs), SC-CO2 explosion, and
ozonolysis), (iii) biological methods (e.g., enzymatic, microbes), and (iv) inhibitors
detoxification methods (Kumar et al., 2022).
Sharma et al. (2020) reported that pretreatment of yard waste via microwaving at
2200 W and 110–200 °C for 2 min increased biogas generation from AD 40.15%.
Basak et al. (2020) observed that hemicellulose, cellulose, and lignin decreased by
62.7, 53.6, and 8.8%, thanks to pretreatment of mixed fruit and vegetable wastes
with Aspergillus fumigatus NITDGPKA3. As a result, biohythane yield and energy
recovery elevated by 53% and 47%, respectively. Another study by Al Ramahi et al.
(2021) revealed that pretreatment of dried dairy sludge via HTC at 210 °C for
30 min boosted CH4 yield to 192% and sludge biodegradability to 30%, while
128 T. A. H. Nguyen et al.

reducing COD by 18%. Matsakas et al. (2019) discovered that using integrated
pretreatments of organic solvent and steam explosion of pruce biomass caused del-
ignification to 79.4% increased cellulose content by 63.3% and saccharification
yield by 61%. Notably, Roque et al. (2019) observed pretreatment of sugar bagasse
with 0.5% H2SO4 at 140 °C for 15 min removing acetic acid 85.4%, phenolics 69%,
formic acid, 5-HMF, and furfural 100%, thereby enhancing inhibitors
detoxification.

5.2 Utilization of Co-Feedstocks

Fang et al. (2017) co-pyrolyzed different blends of paper mill sludge and MSW at
900 °C and found that a mixing ratio of 10% was the best. Similarly, Jung et al.
(2021) revealed that biogas yield obtained with co-digestion of livestock manures
and lignocellulosic biowaste (>0.3 m3/kg VS) was greater than that of single diges-
tion of livestock manures (0.07–0.36 m3/kg VS). Obviously, using co-feedstocks is
an effective method for enhancing biofuel production from biowastes.

5.3 Application of Hybrid Conversion Technologies

Many studies show that using integrated conversion technologies helps intensify
energy generation while reducing adverse environmental impacts. Monlau et al.
(2015) coupled pyrolysis with AD to boost energy recovery from MSW, yielding
bio-oil (58%), biochar (33%), and syngas (9%). Especially, the electricity produc-
tion obtained with the hybrid process elevated by 42% compared to the single
AD. Osman (2020) observed the use of a hybrid process of combustion and pyrolysis
with a DeNOx-catalyzed device reduced NOx emissions by their conversion to H2O
and N2. This strategy can be an excellent choice for MSW that contains constituents
(e.g., food waste) with a high content of nitrogen. Rezaeitavabe et al. (2020) revealed
that the addition of phosphate-laden (10 g/L) biochar in AD of food waste at meso-
philic condition (37 °C) improved biohydrogen production via single-­stage hybrid
dark-photo-fermentation up to 94%. Likewise, Sharma et al. (2020) reported that the
C content and HHV value in the hybrid AD-HTC system (56% and 23 MJ/kg) were
markedly greater than those in the single AD process (44.35% and 16 MJ/kg).

5.4 Tuning of Process Parameters

To enhance biofuel production from biowastes, it is crucial to optimize process


parameters. Salakkam et al. (2019) stated that process parameters may vary, depend-
ing on the selected conversion technology. Yu et al. (2023) employed HTL for
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 129

biofuel production from MSW, reporting that the relationship between temperature
(up to 364.57 °C), reaction time (up to 88.23 min.), and energy yield was propor-
tional, whereas the relationship between waste-to-water ratio and energy yield was
inverse. The highest energy yield was attained at 378 °C, 94.61 min, and waste-to-
water ratio of 0.11 g/mL. Gökçek et al. (2023) employed dark fermentation with
magnetite addition (100 mg/L) to produce biohydrogen from apple pulp waste. The
maximal biohydrogen yield was 73.59 mL/g VS, and biohydrogen production was
elevated by 46%, thanks to intensified bacteria activity.

6 Conclusion and Future Perspectives

Biomass wastes are renewable, biodegradable, abundant, and cheap, making them
promising feedstocks for the production of various biofuels. Depending on the
desired biofuels, different biomass wastes should be rationally utilized.
This review has technically evaluated existing conversion technologies for bio-
fuel production from biomass wastes, highlighting their merits, demerits, and tech-
nical readiness levels. Moreover, the review has critically discussed process
enhancement strategies, including using integrated transformation technologies,
emerging pretreatment processes, co-feedstocks, and process optimization. These
are also potential future research directions.
This review is expected to promote the valorization of biomass wastes as biofuels
in sustainable and cost-effective ways, thereby contributing to (i) preserving nonre-
newable and finite fossil fuels, (ii) generating added values to biomass wastes, (iii)
mitigation of the global climate change, and (iv) promoting the circular economy
toward sustainable development.

References

Al Ramahi, M., Keszthelyi-Szabó, G., & Beszédes, S. (2021). Coupling hydrothermal carboniza-
tion with anaerobic digestion: An evaluation based on energy recovery and hydrochar utiliza-
tion. Biofuel Research Journal, 8(3), 1444–1453.
Amenaghawon, A. N., Evbarunegbe, N. I., & Obahiagbon, K. (2021). Optimum biodiesel produc-
tion from waste vegetable oil using functionalized cow horn catalyst: A comparative evaluation
of some expert systems. Cleaner Engineering and Technology, 4, 100184.
Anekwe, I. M. S., Ar mah, E. K., & Tetteh, E. K. (2022). Bioenergy production: Emerging tech-
nologies. Biomass, Biorefineries and Bioeconomy, 225.
Asadi, N., & Zilouei, H. (2017). Optimization of organosolv pretreatment of rice straw for enhanced
biohydrogen production using Enterobacter aerogenes. Bioresource Technology, 227, 335–344.
Aurnob, A. K., Arnob, A., Kabir, K. B., Islam, M. S., Rahman, M. M., & Kirtania, K. (2022).
Hydrothermal carbonization of biogenic municipal waste for biofuel production. Biomass
Conversion and Biorefinery, 1–9.
Basak, B., Saha, S., Chatterjee, P. K., Ganguly, A., Chang, S. W., & Jeon, B. H. (2020). Pretreatment
of polysaccharidic wastes with cellulolytic Aspergillus fumigatus for enhanced production of
biohythane in a dual-stage process. Bioresource Technology, 299, 122592.
130 T. A. H. Nguyen et al.

Beltrán-Ramírez, F., Orona-Tamayo, D., Cornejo-Corona, I., González-Cervantes, J. L. N., de


Jesús Esparza-Claudio, J., & Quintana-Rodríguez, E. (2019). Agro-industrial waste revalo-
rization: The growing biorefinery. Biomass for Bioenergy-Recent Trends and Future
Challenges, 83–102.
Chandraratne, M. R., & Daful, A. G. (2022). Advances in bioenergy production using fast pyrolysis
and hydrothermal processing (p. 269). Biomass.
Clauser, N. M., González, G., Mendieta, C. M., Kruyeniski, J., Area, M. C., & Vallejos, M. E. (2021).
Biomass waste as sustainable raw material for energy and fuels. Sustainability, 13(2), 794.
Czekała, W., Bartnikowska, S., Dach, J., Janczak, D., Smurzyńska, A., Kozłowski, K., et al. (2018).
The energy value and economic efficiency of solid biofuels produced from digestate and saw-
dust. Energy, 159, 1118–1122.
Edwiges, T., Frare, L., Mayer, B., Lins, L., Triolo, J. M., Flotats, X., & de Mendonça Costa,
M. S. S. (2018). Influence of chemical composition on biochemical methane potential of fruit
and vegetable waste. Waste Management, 71, 618–625.
Fang, S., Yu, Z., Lin, Y., Lin, Y., Fan, Y., Liao, Y., & Ma, X. (2017). A study on experimental
characteristic of co-pyrolysis of municipal solid waste and paper mill sludge with additives.
Applied Thermal Engineering, 111, 292–300.
Garba, A. (2020). Biomass conversion technologies for bioenergy generation: An introduction. In
Biotechnological applications of biomass. IntechOpen.
Gil, A. (2022). Challenges on waste-to-energy for the valorization of industrial wastes: Electricity,
heat and cold, bioliquids and biofuels. Environmental Nanotechnology, Monitoring &
Management, 17, 100615.
Gökçek, Ö. B., Baş, F., Muratçobanoğlu, H., & Demirel, S. (2023). Investigation of the effects
of magnetite addition on biohydrogen production from apple pulp waste. Fuel, 339, 127475.
Gomaa, M. A., & Abed, R. M. (2017). Potential of fecal waste for the production of biomethane,
bioethanol and biodiesel. Journal of Biotechnology, 253, 14–22.
Indrawan, B., Prawisudha, P., & Yoshikawa, K. (2012). Combustion characteristics of chlorine-free
solid fuel produced from municipal solid waste by hydrothermal processing. Energies, 5(11),
4446–4461.
Irmak, S. (2019). Challenges of biomass utilization for biofuels. Biomass for Bioenergy-Recent
Trends and Future Challenges, 1–11.
Jung, S., Shetti, N. P., Reddy, K. R., Nadagouda, M. N., Park, Y. K., Aminabhavi, T. M., & Kwon,
E. E. (2021). Synthesis of different biofuels from livestock waste materials and their potential
as sustainable feedstocks – A review. Energy Conversion and Management, 236, 114038.
Kang, K., Nanda, S., Sun, G., Qiu, L., Gu, Y., Zhang, T., et al. (2019). Microwave-assisted hydro-
thermal carbonization of corn stalk for solid biofuel production: Optimization of process
parameters and characterization of hydrochar. Energy, 186, 115795.
Kim, M., Jung, S., Lee, D. J., Lin, K. Y. A., Jeon, Y. J., Rinklebe, J., et al. (2020). Biodiesel syn-
thesis from swine manure. Bioresource Technology, 317, 124032.
Kumar, R., Kim, T. H., Basak, B., Patil, S. M., Kim, H. H., Ahn, Y., et al. (2022). Emerging
approaches in lignocellulosic biomass pretreatment and anaerobic bioprocesses for sustainable
biofuels production. Journal of Cleaner Production, 333, 130180.
Lee, X. J., Ong, H. C., Gao, W., Ok, Y. S., Chen, W. H., Goh, B. H. H., & Chong, C. T. (2021).
Solid biofuel production from spent coffee ground wastes: Process optimization, characteriza-
tion and kinetic studies. Fuel, 292, 120309.
Lee, J., Kim, S., You, S., & Park, Y. K. (2023). Bioenergy generation from thermochemical conver-
sion of lignocellulosic biomass-based integrated renewable energy systems. Renewable and
Sustainable Energy Reviews, 178, 113240.
Liu, Z., Quek, A., Hoekman, S. K., & Balasubramanian, R. (2013). Production of solid biochar
fuel from waste biomass by hydrothermal carbonization. Fuel, 103, 943–949.
Matsakas, L., Raghavendran, V., Yakimenko, O., Persson, G., Olsson, E., Rova, U., et al. (2019).
Lignin-first biomass fractionation using a hybrid organosolv–steam explosion pretreatment
technology improves the saccharification and fermentability of spruce biomass. Bioresource
Technology, 273, 521–528.
A State of the Art of Biofuel Production Using Biomass Wastes: Future Perspectives 131

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. Applied Energy, 148, 32–38.
Moon, J., Mun, T. Y., Yang, W., Lee, U., Hwang, J., Jang, E., & Choi, C. (2015). Effects of hydro-
thermal treatment of sewage sludge on pyrolysis and steam gasification. Energy Conversion
and Management, 103, 401–407.
Nadir, N., Ismail, N. L., & Hussain, A. S. (2019). Fungal pretreatment of lignocellulosic materials.
In Biomass for bioenergy-recent trends and future challenges. IntechOpen.
Nanda, S., & Berruti, F. (2021). A technical review of bioenergy and resource recovery from
municipal solid waste. Journal of Hazardous Materials, 403, 123970.
Osman, A. I., Deka, T. J., Baruah, D. C., & Rooney, D. W. (2020). Critical challenges in biohydro-
gen production processes from organic feedstocks. Biomass Conversion and Biorefinery, 1–19.
Panjičko, M., Zupančič, G. D., Fanedl, L., Logar, R. M., 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. Journal of Cleaner Production,
166, 519–529.
Rezaeitavabe, F., Saadat, S., Talebbeydokhti, N., Sartaj, M., & Tabatabaei, M. (2020). Enhancing
biohydrogen production from food waste in single-stage hybrid dark-photo fermentation
by addition of two waste materials (exhausted resin and biochar). Biomass and Bioenergy,
143, 105846.
Roque, L. R., Morgado, G. P., Nascimento, V. M., Ienczak, J. L., & Rabelo, S. C. (2019). Liquid-­
liquid extraction: A promising alternative for inhibitors removing of pentoses fermentation.
Fuel, 242, 775–787.
Sahoo, K., Bilek, E., Bergman, R., & Mani, S. (2019). Techno-economic analysis of producing
solid biofuels and biochar from forest residues using portable systems. Applied Energy, 235,
578–590.
Salakkam, A., Plangklang, P., Sittijunda, S., Kongkeitkajorn, M. B., Lunprom, S., & Reungsang,
A. (2019). Biohydrogen and methane production from lignocellulosic materials. Biomass for
Bioenergy-Recent Trends and Future Challenges.
Santana, M. S., Alves, R. P., da Silva Borges, W. M., Francisquini, E., & Guerreiro, M. C. (2020).
Hydrochar production from defective coffee beans by hydrothermal carbonization. Bioresource
Technology, 300, 122653.
Scarlat, N., Motola, V., Dallemand, J. F., Monforti-Ferrario, F., & Mofor, L. (2015). Evaluation of
energy potential of municipal solid waste from African urban areas. Renewable and Sustainable
Energy Reviews, 50, 1269–1286.
Sekoai, P. T., Ayeni, A. O., & Daramola, M. O. (2019). Parametric optimization of biohydrogen
production from potato waste and scale-up study using immobilized anaerobic mixed sludge.
Waste and Biomass Valorization, 10, 1177–1189.
Shafizadeh, A., & Danesh, P. (2022). Biomass and energy production: Thermochemical methods.
Biomass, Biorefineries and Bioeconomy, 247.
Sharma, H. B., Panigrahi, S., Sarmah, A. K., & Dubey, B. K. (2020). Downstream augmentation
of hydrothermal carbonization with anaerobic digestion for integrated biogas and hydrochar
production from the organic fraction of municipal solid waste: A circular economy concept.
Science of the Total Environment, 706, 135907.
Sonu, Rani, G. M., Pathania, D., Umapathi, R., Rustagi, S., Huh, Y. S., Gupta, V. K., et al. (2023).
Agro-waste to sustainable energy: A green strategy of converting agricultural waste to nano-­
enabled energy applications. Science of the Total Environment, 875, 162667.
Tandon, M., Thakur, V., Tiwari, K. L., & Jadhav, S. K. (2018). Enterobacter ludwigii strain
IF2SW-B4 isolated for biohydrogen production from rice bran and de-oiled rice bran.
Environmental Technology & Innovation, 10, 345–354.
U.S. EIA. (2019). EIA releases plant-level U.S. biodiesel production capacity data. U.S. Energy
Information Administration.
132 T. A. H. Nguyen et al.

Wang, L., Li, A., & Chang, Y. (2017). Relationship between enhanced dewaterability and struc-
tural properties of hydrothermal sludge after hydrothermal treatment of excess sludge. Water
Research, 112, 72–82.
Yu, D., Guo, J., Meng, J., & Sun, T. (2023). Biofuel production by hydro-thermal liquefaction of
municipal solid waste: Process characterization and optimization. Chemosphere, 328, 138606.
Zhuang, X., Liu, J., Zhang, Q., Wang, C., Zhan, H., & Ma, L. (2022). A review on the utiliza-
tion of industrial biowaste via hydrothermal carbonization. Renewable and Sustainable Energy
Reviews, 154, 111877.
Role of Pretreatment Approaches
to Generate Value-Added Products Using
Agriculture Biomass

Suman, Deepanshu Awasthi, Nishtha, Nikhil Gakkhar, and Bharat Bajaj

1 Introduction

As a result of the negative effects of using fossil fuels, such as global warming
and greenhouse gas emissions, there is a consistent demand for using alternative
energy sources. To meet the challenges of fossil fuel, this fossil fuel may be
replaced with renewable energy sources. Lignocellulosic biomass (LCB) is con-
sidered the most readily available renewable energy source. Residual wastes and
agricultural biomass are easily available sources of biomass all around the world
(Kumar et al., 2023; Bhatia et al., 2012). Around two billion tons of agro-waste
are produced annually around the world. Agro-waste is a type of waste or by-
products that are produced during various agricultural activities (Dey et al.,
2021). On the basis of its origin, agro-waste is categorized into two types: agro-
residues (comes from agricultural land) and industrial residues (comes from
processing of raw material at industrial sites) (Nath et al., 2023). Agro-residues
and industrial residues can be further subcategorized. Field residue and process-
ing residues are subcategorization of agro-residues. Field residues are waste
materials such as husks, leaves, stalks, and stems that are left on agricultural
fields after harvesting of crops. Process residues are the type of field residues,
which are left behind after the crop is converted into its final product, such as

Suman · Nishtha · B. Bajaj (*)


Centre for Nanoscience & Nanotechnology, Panjab University, Chandigarh, India
e-mail: bharatbajaj@dtu.ac.in
D. Awasthi
Sardar Swaran Singh National Institute of Bio-Energy, Kapurthala, India
N. Gakkhar
Ministry of New and Renewable Energy, Atal Akshay Urja Bhavan, New Delhi, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 133
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_7
134 Suman et al.

cotton linter seeds. The garbage produced by industrial processes includes


potato peel, waste from tea preparation, soybean, and coconut oil cake.
Agro-waste management methods include burning, improper dumping, and add-
ing additives to ruminant and poultry feed (Sadh et al., 2018). Production of aero-
sols such as N2O, CO, CH4, and NOx, as well as a significant loss of soil nutrients
and microbial population, is one consequence of firing of agricultural waste. In
addition, haphazard trash disposal in open spaces causes rotting and related envi-
ronmental problems. Improper waste management can result in contamination of
the land, water, and air, which adds to climate change. Proper use of agricultural
waste has two advantages: It supplies inexpensive biodegradable raw materials, pro-
duces cash and jobs, and mitigates the negative consequences of agro-waste
(Adejumo & Adebiyi, 2020).
The primary components of agro-waste are known as agro- and industrial left-
overs. Hemicellulose, cellulose, and lignin (Fig. 1) are agro- and industrial left-
overs. Sugar polymers are helpful in making the structure of cellulose and
hemicellulose fractions. They are possible sources for fermentation processes, water
filtration, and the creation of adsorbent membranes (Ganguly et al., 2020). Lignin
may be used to produce chemicals, heat, and energy, among other things (Gupta
et al., 2022). Resistance of lignin to enzymatic hydrolysis is a major challenge in
LCB valorization (Banu et al., 2021; Zoghlami & Paës, 2019). Complex assemblies
of component polymers that are naturally resistant to enzymatic conversion produce
difficulties in lignocellulosic biomass valorization. It is required to change the phys-
ical or chemical structure of LCB so that its constituent polymer can be accessed.
To change the physical or chemical structure of LCB, several pretreatment pro-
cesses are required (Yang & Wyman, 2007). Pretreatment processes are helpful to
get over LCB resistance and help to break down LCB into its component parts.
Therefore, for breaking down LCB into its constituent parts, an eco-friendly, effec-
tive, and economical pretreatment method is required. Various chemical, physical,
and biological pretreatment techniques are available for the pretreatment of ligno-
cellulosic biomass (Chundawat et al., 2010; Bułkowska & Klimiuk, 2016). Herein,
this chapter, the use of various pretreatment ways for the derivation of value-added
products from agro-waste biomass, their mechanisms, advantages, and disadvan-
tages are discussed.

Fig. 1 Schematic representation of major components of lignocellulosic biomass


Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 135

2 Components of Agricultural Biomass

The three polymers, that is, cellulose, hemicellulose, and lignin, are responsible for
making up the majority of the biomass originating from agricultural plants (Fig. 1).
These three polymers carry some minor quantities of pectin, protein, extractives,
and ash with them. The proposition or concentration of constituent components may
differ depending on the type of plant (Banu et al., 2021). The proposition of these
components is dependent on various factors such as plant age and plant develop-
mental stages. It is observed that cellulose is more prevalent in hardwoods, but
hemicellulose is more prevalent in wheat straw and leaves. On the basis of kind,
species, and source of the biomass, constituent polymers are bonded to each other
in various matrixes to a varied degree.

2.1 Cellulose

Cellulose is a form of linear macromolecule, which is made from repeating


D-glucopyranose units associated with 1,4-glycosidic linkages (Fig. 2). The annual
output of cellulose as the most prevalent renewable biopolymer on earth is around
1.5 × 1012 tons (Padhi et al., 2023). Hydrogen bonding between intra- and intermo-
lecular molecules of cellulose is responsible for giving characteristics such as
hydrophilicity, chirality, ease of chemical functionalization, insolubility in most sol-
vents, and resistance to degradation. Natural cellulose has both organized (crystal-
line) and disorganized (amorphous) structures within it. The crystallinity of cellulose
ranges from 40% to 70%, which is dependent on the natural source and used extrac-
tion. The amorphous domain in plant structure has lower density and more perme-
ability to chemicals as compared to the crystalline domain. Thus, the amorphous
part of cellulose makes it more malleable. Disordered parts may hydrolyze under
carefully regulated acidic circumstances, while the crystalline regions remain intact.

Fig. 2 Structure of cellulose


136 Suman et al.

2.2 Hemicellulose

Hemicellulose is another prevalent polymer constituent of LCB, that is, hemicellu-


lose, accounts for 20–50% of the LCB. Like cellulose, it is not chemically homog-
enous (Karimi et al., 2013). The branches of hemicellulose consist of short lateral
chains made of various carbohydrates. These carbohydrates include uronic acids
(4-O-methylglucuronic, D-galacturonic acids, and D-glucuronic) and hexoses (glu-
cose, mannose, and galactose), pentoses (xylose, rhamnose, and arabinose,), and
hexoses (glucose, galactose, and mannose). Hemicelluloses consist of brief short
chains linked via beta-(1,4)-glycosidic and beta-(1,3)-glycosidic linkages as shown
in Fig. 3. It can also be partially acetylated; one example of partial acetylation is
heteroxylan. It is found that hemicelluloses have lower molecular weights and fea-
ture branches compared to cellulose due to its little lateral branch that makes it
simple in hydroxylation. Proposition or concentration of hemicellulose is varied
from plant to plant. It has been found that hemicelluloses are majorly made up of
xylan in agricultural biowaste such as straws and grasses, whereas in softwood these
are made up of glucomannan. Most plants contain xylans, which are heteropolysac-
charides having 1,4-linked beta-D-xylopyranose unit backbone chains. Xylan can
also have arabinose, p-coumaric acids, acetic acid, glucuronic acid, and ferulic,
along with xylose. Glucomannan in LCB can be extracted with ease in an alkaline
or acidic environment; xylan can only be retrieved in a more alkaline environment.
The most thermochemically sensitive of the main lignocellulosic constituents are
hemicelluloses. It has been suggested that at least 50% of the hemicellulose in plant
cell walls should be eliminated in order to considerably boost the digestion of cel-
lulose since it is believed that hemicelluloses cover the cellulose fibrils (Banu
et al., 2021).

Fig. 3 Structure of hemicellulose


Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 137

2.3 Lignin

The third most prevalent and most complicated natural polymer in LCB is consid-
ered lignin. Lignin gives structural support, resistance to impact, impermeability,
bending, and compression and protection against microbial assault and oxidative
stress to the cell wall of LCB plants. It is also important for the movement of metab-
olites, nutrients, and water within the plant cell of LCB. As can be seen in Fig. 4, the
lignin is a 3-d network of phenylpropanoid units (p-coumaryl, and sinapyl alcohol
coniferyl) linked by various types of bonding. Monomers of lignin are joined via
alkyl-aryl, aryl-aryl, and alkyl-alkyl and ether linkages. It is observed that varied
kinds of feedstock have different structures and different levels of lignin. Generally,
it has been found that softwoods have the greatest lignin concentrations, whereas
grasses have the lowest. It is determined that lignin from softwood contains more
than 90% coniferyl alcohol, but lignin from hardwood contains coniferyl and sina-
pyl alcohol in various proportions (Karimi et al., 2013). It is well known that lignin
works like an adhesion in LCB that binds the various constituent parts of LCB
together and causes the rendering of LCB constituent solubilization into water. It is
known as a major challenge in the enzymatic and microbiological hydrolysis of
LCB due to its connection to cellulose. As more lignin is removed, biomass digest-
ibility gets improved. The major impacts of lignin are that it physically prevents
enzymes from accessing sugars, and it acts as a potent inhibitor of fermentation

Fig. 4 Structure of lignin


138 Suman et al.

Table 1 Composition of agro-waste from various sources


Agro-waste Cellulose (%) Hemicellulose (%) Lignin (%)
Bamboo 41.8 18 29.3
Corn stalks 50 20 30
Sugarcane bagasse 35 35.8 16.1
Rice husk 37.1 29.4 24.1
Rice straw 35.8 21.5 24.4
Corn stover 38.4 22.9 20.1
Sugarcane tops 43 27 17
Beechwood 44.2 33.5 21.8
Grasses 10–30 25–40 25–50
Barley straw 13.3 34.3 23
Wheat straw 15.6 35 22.3

Fig. 5 Schematic showing various pretreatment methods for the derivation of value-added prod-
ucts from agro-waste biomass

(Banu et al., 2021). Compositions of agro-waste from various agricultural sources


are shown in Table 1.

3 Advantages of Pretreatment of Biomass

Various chemical, physical, and biological pretreatment ways are available for the
pretreatment of LCB as shown in Fig. 5. The major aims of the pretreatment meth-
ods are to lower cellulose crystallinity, to remove hemicellulose and lignin, and to
increase the porosity of extracted LCB material.
Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 139

Pretreatment methods should fulfill some requirements such as it should con-


sume low capital and have low operating costs, and it should be effective on a large
number of LCBs to recover the majority of lignocellulosic components. Pretreatment
methods should prevent by-product production that resists the next fermentation
and hydrolysis processes and the decomposition or loss of carbohydrates. There is
no ideal pretreatment technique available, as there are still bottlenecks in pretreat-
ment methods such as high load of particles and use of high input energy and gen-
eration of inhibitory chemicals such as acids, furans, and phenols. Individual
feedstock needs a different set of pretreatment parameters and mechanism model
that are helpful in designing processes. Pretreatments can be categorized as chemi-
cal, physical, or biological, but there is a good relationship between these processes.
Being one of the most expensive procedures, lignocellulose pretreatment methods
are required to be optimized in particular. In order to increase digestibility, various
thermochemical pretreatment techniques have also been created. There is a signifi-
cant association between the elimination of lignin and hemicellulose and the digest-
ibility of cellulose, as demonstrated by several studies. For the conversion of the
lignin portion of cellulosic biomass, this might negatively impact enzyme hydroly-
sis. The thermochemical pretreatment methods are found to be more useful as com-
pared to biological methods. It works as an energy source and produces possible
by-products that have significant advantages throughout the course of a life cycle
(Bułkowska & Klimiuk, 2016).

3.1 Physical Pretreatment

In order to lower the particle size of biomass, a variety of physical pretreatment


methods such as mechanical devices, including millers, grinders, and screws, as
well as processes such as ultraviolet or microwave radiations, are frequently used
(Mankar et al., 2021). In general, for the physical pretreatment method, mechanical
shearing is used, which takes place at temperatures between 180 and 240 °C (Tu &
Hallett, 2019). As a result, specific surface area, particle sizes, degree of crystallin-
ity, and polymerization vary. In addition, by avoiding the use of chemicals, physical
preparation has the advantage of reducing the generation of waste and inhibitors
(Jȩdrzejczyk et al., 2019). The use of environment-friendly methods including
mechanical, microwave, or ultrasonic pretreatments increases biomass processing
efficiency. Although these techniques typically prevent the synthesis of harmful
compounds, their high-energy usage is still a serious disadvantage (Baruah
et al., 2018).
140 Suman et al.

3.2 Milling

The milling pretreatment method is useful in reducing the size of biomass particle.
There are numerous approaches, such as milling, shredding, and chipping, which
produce biomass with a range of particle sizes. For instance, grinding or milling
biomass often produces particles between 0.2 and 2 mm, whereas chipping typically
results in particle size within 10 and 30 mm. Decreased cellulose polymerization
and cellulose crystallinity and enhanced surface area for enzymatic hydrolysis and
mass transfer due to smaller particle size are but a few of the advantages of milling.
However, there are drawbacks to consider; for example, it demands a significant
amount of energy. Furthermore, lignin is not removed during this process, which
may make it more challenging for the enzymes to hydrolyze cellulose and hemicel-
lulose (Gupta et al., 2022).
A variety of motorized equipment can be used for milling, including wet disk
mills, two-roll mills, ball mills, rod mills, hammer mills, vibratory mills, and two-­
roll mills. Based on the specified milling technique, the duration of processing time,
the kind of biomass, the degree of particle size reduction, and crystallinity can be
determined. In a study of the three milling methods (ball milling, attrition milling,
and planetary milling), it turned out that planetary milling and attrition milling were
more effective at reducing biomass particle size than ball milling. Since it does not
produce any toxic or inhibiting compounds, milling pretreatment is the ideal pre-
liminary pretreatment method for a variety of lignocellulosic feedstock (Gupta
et al., 2022).

3.3 Microwave

During the microwave pretreatment method, LCB is exposed to high-frequency


electromagnetic waves, which typically range from 300 MHz to 300 GHz. The pro-
cess makes use of microwave unique heating capabilities, which instantly produce
localized hotspots inside the biomass and preferentially heat polar molecules such
as water. As a result of the localized heating, the lignocellulosic structure disinte-
grates, which also improves the efficiency of the subsequent enzymatic hydrolysis.
There are two different types of microwave treatment: atmospheric and high pres-
sure. High-pressure microwave pretreatments are conducted in closed reactors at
temperatures ranging from 150 to 250 °C. Rapid and concentrated heating by
microwaves accelerates the decomposition of lignocellulosic biomass by disrupting
the lignin-carbohydrate complex and reducing cellulose crystallinity (Jȩdrzejczyk
et al., 2019). Due to the breakdown of lignocellulosic components, microwave pre-
treatment results in increased yields of fermentable sugars. This improves biocon-
version efficiency overall and reduces the need for expensive enzyme cocktails,
making it particularly beneficial for subsequent fermentation operations (Baruah
et al., 2018). By reducing heat loss to the environment, the selective heating
Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 141

technique employed in microwave pretreatment increases energy efficiency. Despite


encouraging results for microwave pretreatment at the laboratory scale, scaling the
process up to industrial levels creates challenges. The development of durable reac-
tor designs, efficient energy transmission, and uniform microwave heating of enor-
mous volumes of biomass are crucial elements that contribute to unsuccessful
implementation. Commercial-scale microwave pretreatment systems might require
a sizable upfront investment due to the sophisticated equipment and infrastructure
required to handle microwave energy. Cost-cutting initiatives, however, might even-
tually be aided by technological advancements and economies of scale (Peral, 2016).

3.4 Extrusion

Extrusion is a highly effective pretreatment for lignocellulosic biomass that utilizes


mechanical shear forces, elevated temperature, and pressure to disrupt the complex
structure of biomass and enhance its enzymatic digestibility. The biomass feedstock
is compelled through a small die opening during extrusion under precisely con-
trolled conditions, producing heat as a result of viscous dissipation and mechanical
work. The complex lignocellulosic matrix is broken down by the shear and com-
pression forces utilized in extrusion, which results in smaller particles and more
surface area accessible for enzymatic hydrolysis. This improves the extraction of
constituents (Zheng & Rehmann, 2014).
Additionally, extrusion facilitates the removal of hemicelluloses, which are more
readily solubilized by the applied heat and pressure. By removing hemicelluloses,
the porosity of the biomass is increased, which permits greater enzyme penetration
and reduces the amount of enzymes required for efficient hydrolysis. Moreover,
extrusion can result in the partial decrystallization of cellulose, rendering it more
amorphous and susceptible to enzymatic attack. However, extrusion pretreatment
has some drawbacks. The process requires a substantial amount of energy, primarily
in the form of heat, to reach the proper temperature and pressure conditions. The
increased energy demand may increase operating costs and have a detrimental effect
on the environment. In addition, the mechanical forces used during extrusion may
generate inhibitory chemicals, such as furans and phenolics, which can have a del-
eterious effect on subsequent processes. Despite these constraints, pretreating ligno-
cellulosic feedstock such as agricultural leftovers with extrusion has produced
positive results. For instance, research has demonstrated that pretreating maize sto-
ver with extrusion prior to enzymatic digestion can greatly improve enzymatic
digestibility (Wang et al., 2020). Similar to this, it has been discovered that extrud-
ing switch grass improves cellulose hydrolysis and increases ethanol production
(Karunanithy & Muthukumarappan, 2010).
142 Suman et al.

3.5 Chemical Pretreatment

Chemical pretreatment techniques are essential for augmenting the enzymatic


hydrolysis of lignocellulosic biomass, thereby facilitating the conversion of com-
plex polysaccharides into fermentable sugars. Several pretreatment methods are
useful to increase the bioavailability of cellulose and hemicellulose while concur-
rently decreasing lignin content and its inhibitive effects. Liquefied acids, such as
H2SO4 and HCl, are commonly used in hydrolysis of hemicellulose into monomeric
carbohydrates and disrupt the structure of lignin. Alkaline pretreatment, such as
ammonia fiber expansion (AFEX) or sodium hydroxide treatment, modifies the
structure of lignin and promotes hemicellulose removal. Other chemical approaches
include organosolv pretreatment, in which organic solvents such as ethanol or meth-
anol are used to dissolve lignin and facilitate its separation from cellulose (Karimi
et al., 2013).

3.6 Alkaline

The alkali pretreatment method employs potassium, sodium, ammonium, and cal-
cium hydroxides or anhydrous ammonia for biomass pretreatment (Xu & Sun,
2016). During this pretreatment method, saponification and salvation reactions fol-
lowed by the elimination of acetyl and uronic acid groups in the hemicellulose take
place (Suman et al., 2023). Consequently, enzyme accessibility to the cellulose frac-
tion is enhanced, which facilitates the fermentation of carbohydrates for ethanol
production. These reactions increase the cellulose internal surface area while
decreasing its crystallinity and degree of polymerization. Various parameters,
including alkali concentration and type, temperature, reaction duration, and bio-
mass characteristics, determine the efficacy of alkali pretreatment (Loow et al.,
2016). For the conversion of lignocellulosic biomass, alkaline pretreatment has
numerous benefits. For example, it efficiently eliminates lignin, reducing biomass
recalcitrance and enhancing the efficiency of enzymatic hydrolysis. Additionally, it
has the potential to partially hydrolyze hemicellulose into soluble carbohydrates,
which can then be used to produce bioethanol. Furthermore, alkaline pretreatment
facilitates the fractionation of biomass components, allowing the extraction of high-­
value lignin derivatives for a variety of applications.
However, there are assortments of disadvantages to alkaline pretreatment. In
addition, the use of alkaline reagents may generate waste streams containing high
concentrations of lignin and alkali, which must be appropriately treated. In addition,
the procedure may result in cellulose degradation, leading to sucrose loss and a
decrease in overall process efficiency (Xu & Sun, 2016).
Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 143

3.7 Bleaching

Bleaching is an oxidative pretreatment procedure that treats lignocellulose materials


with an oxidizing chemical such as hydrogen peroxide. The removal of hemicellu-
lose and lignin increases cellulose exposure and accessibility, thereby facilitating
the enzymatic or chemical conversion of cellulose into fermentable carbohydrates
or other desirable substances (Dutra et al., 2018). It is found that the elimination of
hemicellulose reduces the formation of inhibitory metabolites during subsequent
fermentation processes. Biomass oxidation is not a selective process. In the case of
lignin extraction, there is always a loss of cellulose and hemicellulose. In the pres-
ence of an oxidizing agent, the oxidation of aromatic rings to carboxylic acids
makes the delignification process effective. Other delignification reactions occur
alongside including electrophilic substitution, alkyl-aryl, and side chain displace-
ment bond cleavage. The bleaching of lignocellulosic biomass increases its enzy-
matic digestibility, resulting in greater sugar yields. Additionally, bleaching
eliminates contaminants and colored compounds, resulting in biomass that is lighter
and more aesthetically appealing. This is notably important in industries that use
biomass as a raw material, such as pulp and paper production, where the final prod-
uct color is significant. However, bleaching has certain disadvantages. The process
can be energy-intensive, necessitating high temperatures and pressures, and it may
involve the use of harsh compounds that are environmentally hazardous (Su
et al., 2015).

3.8 Acid Hydrolysis

Using acid catalysts, an acid hydrolysis pretreatment procedure is performed for


lignocellulosic biomass that degrades complex polysaccharides into simple sugars.
Concentrated acids such as H2SO4 and HCl are used in acid hydrolysis pretreatment
of biomass. Concentrated acids, despite being potent cellulose hydrolysis agents,
are toxic and corrosive hazardous, thus requiring corrosion-proof reactors. The con-
centrated acid must be procured after hydrolysis to make the procedure economi-
cally viable. Dilute acid hydrolysis has been effectively substituted for concentrated
acid hydrolysis in the pretreatment of biomass. It has been applied to numerous
feedstock types, including softwood, herbaceous crops, hardwood, agricultural resi-
dues, municipal solid waste, and wastepaper. It was effective for the vast number of
biomass pretreatment. Typically, two different ways are known for dilute acid pre-
treatment. The first way is high-temperature pretreatment with regular flow methods
with a low biomass concentration. The second way is the use of low-temperature
batch processing methods in a high biomass concentration (Singh et al., 2016;
Mathew et al., 2016). Neutralization of pH is needed for subsequent enzymatic
hydrolysis. The acid promotes hemicellulose hydrolysis resulting in the formation
144 Suman et al.

of monomeric carbohydrates, primarily xylose, glucose, and arabinose. Additionally,


the acid degrades lignin, making cellulose more accessible to enzymes.
As a pretreatment technique for lignocellulosic biomass, acid hydrolysis has
numerous advantages. It effectively solubilizes hemicellulose and degrades lignin,
resulting in enhanced enzymatic digestibility of cellulose. Consequently, compared
to other pretreatment procedures, it can be conducted under extremely mild condi-
tions, thereby reducing energy demands. In addition, acid hydrolysis has a high rate
of sugar recovery, making it an efficient method for producing sugar. However, the
acid used in the procedure may inhibit subsequent enzymatic hydrolysis or fermen-
tation stages, necessitating neutralization and cleansing. In addition, acid hydrolysis
can generate metabolites such as furfural and hydroxymethylfurfural, which may
impede fermentation or have economic repercussions on subsequent operations
(Padhi et al., 2023).

3.9 Ionic Liquids

Ionic liquids (ILs) have emerged as a viable pretreatment method for LCB due to
their distinctive properties and ability to dissolve and depolymerize biomass com-
ponents efficiently. Composed of bulky organic cations and inorganic or organic
anions, ILs are salts with melting points below 100 °C. Their modifiable character-
istics, including low vapor pressure, exceptional thermal stability, and high biomass
solubility, make them desirable for biomass pretreatment. ILs can disrupt the struc-
ture of lignocellulosic materials by disrupting the hydrogen bonding within con-
stituent polymers of LCB, thereby facilitating subsequent hydrolysis or fermentation
processes (Mosier et al., 2005). The interaction between ILs and biomass compo-
nents helps us comprehend the breakdown of biomass in ILs. Cellulose, for exam-
ple, generates substantial hydrogen bonding and cation interactions with ILs,
whereas hemicellulose and lignin have weaker interactions. To achieve optimal bio-
mass fractionation and maximize sugar yields, it is necessary to customize ILs and
process variables such as duration and temperature. In addition, since ILs are recy-
clable and reusable, they are economically viable for large-scale applications.
Several studies have demonstrated that ILs substantially increase sugar yields and
enzymatic digestibility of pretreated biomass, indicating their potential as an effi-
cient pretreatment strategy for lignocellulosic biomass (Binder & Raines, 2009).
For instance, 1-ethyl-3-methylimidazolium acetate has been widely used as an IL
for biomass pretreatment, exhibiting effective delignification and enhanced enzy-
matic hydrolysis (Sun et al., 2009). Other ILs, including cholinium-based ILs and
protic ILs, have been investigated for biomass pretreatment (Brandt et al., 2011).
However, obstacles persist in the implementation of IL-based pretreatment methods
on a large scale. Key considerations for IL-based pretreatment include the high cost
of IL production and energy-intensive recovery processes (Li et al., 2010).
Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 145

3.10 Deep Eutectic Solvents

Deep eutectic solvents (DESs) are known as viable pretreatment strategy for LCB
due to owing their properties such as its unique properties to solubilize biomass
constituents efficiently (Liu et al., 2019). DESs are combinations of H2-donating
and H2-accepting entities. Interactions between hydrogen-donating and hydrogen-­
accepting entities present in DES aid in the degradation of LCB and enable the
extraction of its constituent components. Choline chloride and urea are used for the
pretreatment of LCB, which is one example of a commonly used DES. Urea serves
as the HBA, and choline chloride serves as the HBD in this DES. The DES effec-
tively breaks down the structure of the biomass, allowing for the dissolution of
constituent components. This pretreatment method has improved the accessibility
of cellulose for subsequent enzymatic hydrolysis by eliminating lignin from bio-
mass. The wheat straw pretreatment with choline chloride and glycerol DES pro-
duced a high lignin removal efficiency of 97.5% and a cellulose conversion of
78.9% (Xing et al., 2018). Since DESs are composed of non-toxic and biodegrad-
able components, they are generally considered to be environmentally friendly. In
addition, DES is simple to synthesize and can be modified to specific biomass com-
positions, allowing for a customized strategy for different substrates (Liu et al., 2019).

3.11 Organosolv Pretreatment

Organosolv pretreatment is a way of enhancing the enzymatic digestibility of LCB


for the production of biofuels. This process utilizes a combination of organic sol-
vents and acids to deconstruct the complex structure of lignocellulose. Solvents
such as methanol, ethanol, and acetone disrupt hydrogen bonds within biomass,
causing lignin to separate from cellulose and hemicellulose. The organosolv pre-
treatment process consists of three major steps: impregnation, hydrolysis, and
recovery. During the impregnation procedure, the biomass is soaked in the organic
solvent to enhance its availability to the solvent and facilitate further reactions.
During hydrolysis, an acid catalyst is applied to the solvent-treated biomass to con-
vert hemicellulose into soluble carbohydrates. Lignin is partially dissolved and iso-
lated from the cellulose. During the recovery phase, the solvent is separated from
the solid residue, and both can be further processed and reused.
Pretreatment with organosolv has numerous advantages. First, it eliminates lig-
nin, which has been depicted to inhibit enzymatic hydrolysis. Reduced lignin con-
tent enhances cellulose enzyme accessibility, resulting in increased sugar yields.
Second, the hemicellulose fraction produced during the process is usable as a valu-
able material. Besides, compared to harsher chemical treatments, the organic sol-
vents used in organosolv pretreatment are more eco-friendly, making it a more
sustainable choice. However, organosolv pretreatment has a number of demerits.
The use of acids and organic solvents makes this method relatively expensive. In
146 Suman et al.

addition, recovering these solvents can be difficult and energy-intensive, which


increases the overall expenditure. Various biomass feedstocks may require different
pretreatment conditions, reducing the global applicability of the process.
Multiple studies have demonstrated that organosolv pretreatment is effective. For
instance, researchers have successfully used ethanol-water mixtures as solvents in
the processing of lignocellulosic materials such as wheat straw, corn stover, and
switch grass. They removed a substantial amount of lignin and improved enzymatic
digestibility, resulting in improved sugar yields for subsequent biofuel production.
Enhancing the enzymatic digestibility of LCB by means of organosolv pretreatment
appears to be a viable option. While it has benefits such as efficient lignin removal
and the potential to recover products with added value, it also has issues with cost
and solvent recovery (Jȩdrzejczyk et al., 2019; Kant et al., 2020). Further research
is required to optimize process conditions and devise cost-effective solutions for
widespread adoption.

3.12 Biological Pretreatment

Biological pretreatment is another promising way for the pretreatment of LCB,


which enhances LCB conversion into biofuels and other bio-based compounds.
Primarily, microorganisms such as fungi and bacteria are used in this method to
break down complex structures of LCB into simpler constituents. The main focus is
to break lignin, which acts as a protective layer, as well as hemicellulose, which is
more accessible than cellulose but still requires enzymatic hydrolysis for complete
breakdown (Amin et al., 2017).
The employing fungi involving Pleurotus ostreatus, Phanerochaete chrysospo-
rium, and Trametes versicolor are examples of biological pretreatment of
LCB. These fungi produce ligninolytic enzymes such as lignin peroxidases, manga-
nese peroxidases, and laccases, which can break down lignin efficiently. The fungi
produce free radicals during pretreatment, which split the complex lignin polymer
into smaller, more manageable bits (Alvira et al., 2010).
Biological pretreatment is environmentally benign because it functions under
mild settings and eliminates the use of harsh chemicals that are commonly used in
other pretreatment procedures. Furthermore, because the process is very unique to
lignin breakdown, the cellulose structure is preserved. Furthermore, biological pre-
treatment can be performed utilizing both native and genetically engineered micro-
organisms, allowing for strain optimization and tailoring specific biocatalysts to
target certain biomass compositions.
However, biological pretreatment possesses a few challenges. The process is
relatively slow and may be influenced by environmental conditions such as moisture
content, temperature, and pH, all of which can impact microbial activity.
Furthermore, the lack of a standardized technique and the necessity for optimization
to varied feedstock can stymie its widespread adoption.
Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 147

3.13 Physicochemical Pretreatment

3.13.1 Steam Explosion

Steam explosion is a highly effective LCB pretreatment technology that is widely


employed in biofuel production and other applications (Ziegler-Devin et al., 2021).
LCB is exposed to high-pressure steam and then rapid depressurization, causing
physical disruption of its structure. The initial steam treatment softens the lignocel-
lulosic matrix, making the biomass components more accessible. As a result of the
swift pressure release, the dissolved steam rapidly expands, causing mechanical
forces that cause the biomass to physically break down and disrupt. The harshness
of the steam explosion is able to be adjusted by modifying parameters such as tem-
perature, pressure, and residence duration, allowing for customization based on
individual biomass feedstock and desired outputs. The capacity of steam explosion
to break down lignin, a complex and stiff component of biomass that impedes enzy-
matic hydrolysis, is one of its primary advantages. The breakdown of lignin struc-
ture during steam explosion makes cellulose and hemicellulose more accessible to
enzymatic degradation, increasing the subsequent enzymatic saccharification pro-
cess (Ziegler-Devin et al., 2021). This increases sugar yields, which can then be
fermented into biofuels or used to make other value-added products. Additionally,
steam explosion aids in the removal of extractives and hemicellulose, minimizing
the production of inhibitory compounds during downstream processes.
Agricultural wastes, dedicated energy crops, and forest residue have all been suc-
cessfully steam exploded as lignocellulosic biomass feedstock (Auxenfans et al.,
2017). Steam explosion pretreatment has been shown in studies to significantly
boost sugar release and overall bioconversion efficiency. For example, depending
on the feedstock and process conditions, researchers have documented significant
increases in enzymatic digestibility of lignocellulosic biomass ranging from 30% to
90%. The ability of steam explosion to fractionate LCB into its constituent compo-
nents is one of its key advantages. This promotes the subsequent enzymatic hydro-
lysis of cellulose to glucose for the manufacture of bioethanol. By raising the surface
area of cellulose and decreasing its crystallinity, steam explosion improves its
accessibility to enzymes. Furthermore, this pretreatment approach can make the
lignin structure more susceptible to enzymatic degradation, allowing for more effec-
tive lignin valorization. However, there are some drawbacks associated with steam
explosion. High-energy inputs are required to generate steam and maintain the cor-
rect pressure, resulting in greater operational expenses. Furthermore, the harshness
of the treatment may result in the development of unwanted compounds such as
hydroxymethylfurfural and furfural, which can disrupt future fermentation pro-
cesses. To reduce these by-products, careful process optimization is required. To
summarize, steam explosion is a good pretreatment way for LCB, providing bene-
fits such as component fractionation, greater enzyme accessibility, and increased
lignin valorization. Despite the related energy expenditures and inhibitory com-
pound production, careful optimization can offset these drawbacks. Steam
148 Suman et al.

explosion has significant potential for converting LCB into biofuels and other value-
added products (Lu et al., 2002).

3.13.2 Ammonia Fiber Explosion

The ammonia fiber explosion (AFEX) technique includes treating LCB with ammo-
nia at high temperatures, prompting the lignocellulosic structure to be disrupted.
AFEX exposes LCB to liquid ammonia at high pressure and then rapidly releases it,
resulting in a hotheaded decompression. A quick decrease in pressure results
in localized heating and swelling, causing the biomass structure to be disrupted and
constituent components to be more accessible. Ammonia serves as a catalyst, assist-
ing in the rupture of lignin structure, which is a complex and inflexible component
of biomass. AFEX capacity to successfully pretreat a widespread range of LCB
feedstock, allowing the synthesis of fermentable sugars for biofuel and biochemical
production, is one of its primary benefits. AFEX can produce high sugar yields,
while reducing the requirement for expensive enzyme additions in future hydrolysis
procedures (Nieder-Heitmann et al., 2020). Furthermore, this approach effectively
preserves the biomass cellulose and hemicellulose fractions, making them more
readily available for downstream processing processes.
Despite its benefits, AFEX has significant limitations. Ammonia use can end up
with the development of dangerous chemicals such as ammonia nitrites and amines,
which must be handled correctly during the process. Furthermore, due to the high
temperatures and pressures involved, AFEX necessitates considerable energy
inputs. The operational costs associated with ammonia handling and recovery can
also be a challenge. Several studies have demonstrated the effectiveness of
AFEX. For instance, it was found that AFEX pretreatment of corn stover improved
enzymatic hydrolysis efficiency, resulting in higher sugar yields. Other studies have
explored the application of AFEX on various lignocellulosic feedstock, including
switch grass, sugarcane bagasse, and agricultural residues, with positive results
(Chundawat et al., 2010).
In conclusion, AFEX is a promising pretreatment method for LCB. It offers
advantages such as enhanced approachability of LCB constituent components,
improved yield of sugar, and versatility with different feedstock. However, chal-
lenges related to ammonia handling, energy consumption, and the potential forma-
tion of harmful compounds need to be addressed for its widespread adoption.

3.14 Liquid Hot Water

This pretreatment method is a thermochemical pretreatment method, which com-


prises treating LCB with H2O at high temperatures and pressures ranging from
160 to 250 °C and 1 to 25 MPa. Because of its efficiency in disturbing the com-
plex structure of LCB and increasing its enzymatic digestibility, this
Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 149

pretreatment approach has received a lot of interest in the area of LCB conver-
sion (Zhuang et al., 2016).
Water functions as a catalyst and solvent during LHW pretreatment, accelerating
the breakdown of lignocellulosic components. The high pressure and temperature
and pressure result in hemicellulose depolymerization, lignin elimination, and par-
tial hydrolysis of cellulose. As a result, the biomass is more amenable to enzymatic
hydrolysis, giving good sugar yields.
One of the primary benefits of LHW pretreatment is that it does not need the use
of chemicals, making it an environmentally friendly solution. Furthermore, LHW
pretreatment can be accomplished with water as the sole solvent, reducing the
expense of other solvents. The method is also quite adaptable, as it may be used
with a wide variety of lignocellulosic feedstock, including agriculture wastes, spe-
cific energy crops, and forestry residues (Li et al., 2022).
However, there are certain drawbacks to using LHW pretreatment. The develop-
ment of undesired chemicals such as hydroxymethylfurfural and furfural during
pretreatment results in resistance in further enzymatic hydrolysis and fermentation
steps (Tan et al., 2021). Despite challenges, LHW pretreatment has shown encour-
aging effects in a number of trials. Researchers, for example, have observed consid-
erable gains in enzymatic digestibility and sugar yields when treating diverse
feedstock with LHW. These findings emphasize LHW pretreatment potential as an
effective method of biomass conversion.

4 Conclusion and Future Perspectives

Various chemical, physical, and biological pretreatment methods are available for
the pretreatment of LCB. It has been found that pretreatment of LCB can change
natural properties of biomass. The degree of LCB degradation and its morphologi-
cal characteristics are different in each pretreatment method. By valuation of vari-
ous methods, it has been observed that there is a need to find cost-effective, efficient,
and environment-friendly, which can result in the complete degradation of LCB into
value-added materials. Furthermore, there is a need for future work to understand
the mechanism of pretreatment reaction so that appropriate method can be made.
Additionally, the analysis suggests that utilizing a combination of physicochemical
and thermochemical degradation methods may offer a more efficient approach for
delignifying biomass.

References

Adejumo, I. O., & Adebiyi, O. (2020). Agricultural solid wastes: Causes, effects, and effec-
tive management. Intech, 11(tourism), 13. Available at: https://www.intechopen.com/books/
advanced-­biometric-­technologies/liveness-­detection-­in-­biometrics
150 Suman et al.

Alvira, P., et al. (2010). Pretreatment technologies for an efficient bioethanol production pro-
cess based on enzymatic hydrolysis: A review. Bioresource Technology, 101(13), 4851–4861.
Available at: https://doi.org/10.1016/j.biortech.2009.11.093
Amin, F. R., et al. (2017). Pretreatment methods of lignocellulosic biomass for anaerobic diges-
tion. AMB Express, 7(1). Available at: https://doi.org/10.1186/s13568-­017-­0375-­4
Auxenfans, T., et al. (2017). Understanding the structural and chemical changes of plant biomass
following steam explosion pretreatment. Biotechnology for Biofuels, 10(1), 1–16. Available at:
https://doi.org/10.1186/s13068-­017-­0718-­z
Banu, J., R., et al. (2021). Lignocellulosic biomass pretreatment for enhanced bioenergy recov-
ery: Effect of lignocelluloses recalcitrance and enhancement strategies. Frontiers in Energy
Research, 9(November), 1–17. Available at: https://doi.org/10.3389/fenrg.2021.646057
Baruah, J., et al. (2018). Recent trends in the pretreatment of lignocellulosic biomass for value-­
added products. Frontiers in Energy Research, 6(DEC), 1–19. Available at: https://doi.
org/10.3389/fenrg.2018.00141
Bhatia, L., Johri, S., & Ahmad, R. (2012). An economic and ecological perspective of ethanol
production from renewable agro waste: A review. AMB Express, 2(1), 1–19.
Binder, J. B., & Raines, R. T. (2009). Simple chemical transformation of lignocellulosic bio-
mass into furans for fuels and chemicals. Journal of the American Chemical Society, 131(5),
1979–1985. Available at: https://doi.org/10.1021/ja808537j
Brandt, A., et al. (2011). Ionic liquid pretreatment of lignocellulosic biomass with ionic liquid–
water mixtures. Green Chemistry, 13(9), 2489–2499. Available at: https://doi.org/10.1039/
c1gc15374a
Bułkowska, K., & Klimiuk, E. (2016). Pretreatment of lignocellulosic biomass. In Biomass for
biofuels. Available at: https://doi.org/10.1201/9781315226422
Chundawat, S. P. S., et al. (2010). Multifaceted characterization of cell wall decomposition
products formed during ammonia fiber expansion (AFEX) and dilute acid based pretreat-
ments. Bioresource Technology, 101(21), 8429–8438. Available at: https://doi.org/10.1016/j.
biortech.2010.06.027
Dey, T., et al. (2021). Valorization of agro-waste into value added products for sustainable devel-
opment. Bioresource Technology Reports, 16(August), 100834. Available at: https://doi.
org/10.1016/j.biteb.2021.100834
Dutra, E. D., et al. (2018). Alkaline hydrogen peroxide pretreatment of lignocellulosic biomass:
Status and perspectives. Biomass Conversion and Biorefinery, 8(1), 225–234. Available at:
https://doi.org/10.1007/s13399-­017-­0277-­3
Ganguly, P., et al. (2020). Valorization of food waste: Extraction of cellulose, lignin and their
application in energy use and water treatment. Fuel, 280(July), 118581. Available at: https://
doi.org/10.1016/j.fuel.2020.118581
Gupta, N., et al. (2022). Biomass conversion of agricultural waste residues for different appli-
cations: A comprehensive review. Environmental Science and Pollution Research, 29(49),
73622–73647. Available at: https://doi.org/10.1007/s11356-­022-­22802-­6
Jȩdrzejczyk, M., et al. (2019). Physical and chemical pretreatment of lignocellulosic biomass. In
Second and third generation of feedstocks: The evolution of biofuels. Available at: https://doi.
org/10.1016/B978-­0-­12-­815162-­4.00006-­9
Kant, S., et al. (2020). Recent developments in pretreatment technologies on lignocellu-
losic biomass: Effect of key parameters, technological improvements, and challenges.
Bioresource Technology, 300(October 2019), 122724. Available at: https://doi.org/10.1016/j.
biortech.2019.122724
Karimi, K., Shafiei, M., & Kumar, R. (2013). Progress in physical and chemical pretreat-
ment of lignocellulosic biomass. In Biofuel Technologies. Available at: https://doi.
org/10.1007/978-­3-­642-­34519-­7
Karunanithy, C., & Muthukumarappan, K. (2010). Effect of extruder parameters and moisture con-
tent of switchgrass, prairie cord grass on sugar recovery from enzymatic hydrolysis. Applied
Role of Pretreatment Approaches to Generate Value-Added Products Using Agriculture… 151

Biochemistry and Biotechnology, 162(6), 1785–1803. Available at: https://doi.org/10.1007/


s12010-­010-­8959-­3
Kumar, P., et al. (2023). Utilization of agricultural waste biomass and recycling toward circu-
lar bioeconomy United State of America. Environmental Science and Pollution Research,
8526–8539. Available at: https://doi.org/10.1007/s11356-­022-­20669-­1
Li, C., et al. (2010). Comparison of dilute acid and ionic liquid pretreatment of switchgrass:
Biomass recalcitrance, delignification and enzymatic saccharification. Bioresource Technology,
101(13), 4900–4906. Available at: https://doi.org/10.1016/j.biortech.2009.10.066
Li, X., et al. (2022). Improving enzymatic hydrolysis of lignocellulosic biomass by bio-­coordinated
physicochemical pretreatment—A review. Energy Reports, 8, 696–709. Available at: https://
doi.org/10.1016/j.egyr.2021.12.015
Liu, Y., et al. (2019). Enhanced enzymatic hydrolysis and lignin extraction of wheat straw by
triethylbenzyl ammonium chloride/lactic acid-based deep eutectic solvent pretreatment. ACS
Omega, 4(22), 19829–19839. Available at: https://doi.org/10.1021/acsomega.9b02709
Loow, Y. L., et al. (2016). Typical conversion of lignocellulosic biomass into reducing sugars using
dilute acid hydrolysis and alkaline pretreatment. Cellulose, 23(3), 1491–1520. Available at:
https://doi.org/10.1007/s10570-­016-­0936-­8
Lu, Y., et al. (2002). Cellulase adsorption and an evaluation of enzyme recycle during hydrolysis of
steam-exploded softwood residues. Applied Biochemistry and Biotechnology – Part A Enzyme
Engineering and Biotechnology, 98–100, 641–654. Available at: https://doi.org/10.1385/
ABAB:98-­100:1-­9:641
Mankar, A. R., et al. (2021). Pretreatment of lignocellulosic biomass: A review on recent
advances. Bioresource Technology, 334(April), 125235. Available at: https://doi.org/10.1016/j.
biortech.2021.125235
Mathew, A. K., et al. (2016). An evaluation of dilute acid and ammonia fiber explosion pretreat-
ment for cellulosic ethanol production. Bioresource Technology, 199, 13–20. Available at:
https://doi.org/10.1016/j.biortech.2015.08.121
Mosier, N., et al. (2005). Features of promising technologies for pretreatment of lignocellulosic
biomass. Bioresource Technology, 96(6), 673–686. Available at: https://doi.org/10.1016/j.
biortech.2004.06.025
Nath, P. C., et al. (2023). Biogeneration of valuable nanomaterials from agro-wastes: A compre-
hensive review. Agronomy, 13(2). Available at: https://doi.org/10.3390/agronomy13020561
Nieder-Heitmann, M., et al. (2020). Economic evaluation and comparison of succinic acid and
electricity co-production from sugarcane bagasse and trash lignocelluloses in a biorefinery,
using different pretreatment methods: Dilute acid (H2SO4), alkaline (NaOH), organosolv,
ammonia fibre expa. Biofuels, Bioproducts and Biorefining, 14(1), 55–77. Available at: https://
doi.org/10.1002/bbb.2020
Padhi, S., et al. (2023). Nanocellulose from agro-waste: A comprehensive review of extraction
methods and applications. Reviews in Environmental Science and Biotechnology, 22(1), 1–27.
Available at: https://doi.org/10.1007/s11157-­023-­09643-­6
Peral, C. (2016). Biomass pretreatment strategies (technologies, environmental performance, eco-
nomic considerations, industrial implementation). In Biotransformation of agricultural waste
and by-products: The food, feed, fibre, fuel (4F) economy. Elsevier Inc. Available at: https://doi.
org/10.1016/B978-­0-­12-­803622-­8.00005-­7
Sadh, P. K., et al. (2018). Agro-industrial wastes and their utilization using solid state fermentation:
A review. Bioresources and Bioprocessing, 5(1), 1–15. Available at: https://doi.org/10.1186/
s40643-­017-­0187-­z
Singh, J., et al. (2016). Different pretreatment methods of lignocellulosic biomass for use in bio-
fuel production. Intech, 11(tourism), 13. Available at: https://www.intechopen.com/books/
advanced-­biometric-­technologies/liveness-­detection-­in-­biometrics
Su, Y., et al. (2015). Fractional pretreatment of lignocellulose by alkaline hydrogen peroxide:
Characterization of its major components. Food and Bioproducts Processing, 94(February),
322–330. Available at: https://doi.org/10.1016/j.fbp.2014.04.001
152 Suman et al.

Suman, et al. (2023). Single step heating for facile extraction of cellulose fibers from rice straw and
its copper oxide nanoparticles coating for catalytic reduction application. Waste and Biomass
Valorization, (0123456789), 1–15. Available at: https://doi.org/10.1007/s12649-­023-­02203-­7
Sun, N., et al. (2009). Complete dissolution and partial delignification of wood in the ionic liquid
1-ethyl-3-methylimidazolium acetate. Green Chemistry, 11(5), 646–665. Available at: https://
doi.org/10.1039/b822702k
Tan, Z., et al. (2021). Inhibition and disinhibition of 5-hydroxymethylfurfural in anaerobic fermen-
tation: A review. Chemical Engineering Journal, 424(April), 130560. Available at: https://doi.
org/10.1016/j.cej.2021.130560
Tu, W. C., & Hallett, J. P. (2019). Recent advances in the pretreatment of lignocellulosic biomass.
Current Opinion in Green and Sustainable Chemistry, 20, 11–17. Available at: https://doi.
org/10.1016/j.cogsc.2019.07.004
Wang, Z., et al. (2020). Enhancing enzymatic hydrolysis of corn stover by twin-screw extrusion
pretreatment. Industrial Crops and Products, 143(October 2019), 111960. Available at: https://
doi.org/10.1016/j.indcrop.2019.111960
Xing, W., et al. (2018). Novel dihydrogen-bonding deep eutectic solvents: Pretreatment of rice
straw for butanol fermentation featuring enzyme recycling and high solvent yield. Chemical
Engineering Journal, 333(September 2017), 712–720. Available at: https://doi.org/10.1016/j.
cej.2017.09.176
Xu, J. K., & Sun, R. C. (2016). Recent advances in alkaline pretreatment of lignocellulosic bio-
mass. In Biomass fractionation technologies for a lignocellulosic feedstock based biorefinery
(pp. 431–459). Available at: https://doi.org/10.1016/B978-­0-­12-­802323-­5.00019-­0
Yang, B., & Wyman, C. E. (2007). Pretreatment: The key to unlocking low-cost cellulosic ethanol.
Biofuels, Bioproducts and Biorefining, 2, 26–40. Available at: https://doi.org/10.1002/BBB
Zheng, J., & Rehmann, L. (2014). Extrusion pretreatment of lignocellulosic biomass: A review.
International Journal of Molecular Sciences, 15(10), 18967–18984. Available at: https://doi.
org/10.3390/ijms151018967
Zhuang, X., et al. (2016). Liquid hot water pretreatment of lignocellulosic biomass for bioethanol
production accompanying with high valuable products. Bioresource Technology, 199, 68–75.
Available at: https://doi.org/10.1016/j.biortech.2015.08.051
Ziegler-Devin, I., et al. (2021). Steam explosion pretreatment of lignocellulosic biomass: A mini-­
review of theoretical and experimental approaches. Frontiers in Chemistry, 9(November), 1–7.
Available at: https://doi.org/10.3389/fchem.2021.705358
Zoghlami, A., & Paës, G. (2019). Lignocellulosic biomass: Understanding recalcitrance and pre-
dicting hydrolysis, 7(December). Available at: https://doi.org/10.3389/fchem.2019.00874
Utilising Biomass-Derived Composites
in 3D Printing to Develop Eco-Friendly
Environment

Chetan Chauhan, Varsha Rani, Mukesh Kumar, and Rishubh Motla

1 Introduction

In recent times, the manufacturing sector has undergone a significant shift due to the
advent of three-dimensional (3D) printing, also known as additive manufacturing
(AM) technology. The utilisation of this technology has brought about a significant
transformation in the supply chain industry, as it allows for the implementation of
localised and even personalised manufacturing processes for goods (Ji et al., 2020).
The classification of the additive manufacturing (AM) process, as outlined in ASTM
F2792-12a, involves categorising it into seven distinct groups according to the lay-
ering mechanism employed (Wang et al., 2017). Among these groups, material
extrusion (ME) is considered the most favourable method due to its cost-­
effectiveness, ease of setup, and extensive selection of materials. The utilisation of
ME machines, specifically those designed for desktop use, has experienced a surge
in popularity due to their cost-effectiveness and capacity to enable users to indepen-
dently manufacture their own essential items and consumer goods through a do-it-­
yourself (DIY) approach (Astm, 2015).
The utilisation of filament-formed materials is a common practice in the field of
ME machines. These materials are typically composed of thermoplastic polymers,
such as polylactic acid (PLA) and acrylonitrile butadiene styrene (ABS).
Nevertheless, the growing utilisation of 3D printing for the production of end-use
products in various sectors has led to heightened apprehensions about the environ-
mental implications associated with the disposal of conventional plastic materials
and the potential risks they pose in terms of toxicity. In order to tackle these

C. Chauhan (*) · M. Kumar · R. Motla


Department of Floriculture and Landscaping Architecture, Sardar Vallabhbhai Patel
University of Agriculture & Technology, Meerut, India
V. Rani
Department of Agriculture, Meerut Institute of Technology, Meerut, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 153
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_8
154 C. Chauhan et al.

concerns, the development of bio-based plastic materials has been undertaken as a


means to decrease dependence on petrochemical polymers (Ligon et al., 2017).
PLA is a prominent example of a bio-based polymer that is obtained from the fer-
mentation of plant materials. It possesses satisfactory mechanical strength, although
its thermal properties are somewhat restricted. Bio-based polyethylene (PE) is a
frequently utilised material that demonstrates applicability in the production of con-
sumer goods. However, its mechanical performance is comparatively lower when
employed in industrial settings. The development of bio-based polycarbonate (PC)
has emerged as a potential alternative for engineering plastic materials, aiming to
replace the conventional fossil-fuel-based PC. This substitution is motivated by
concerns regarding the potential toxicity associated with the latter. The utilisation of
bio-based polycarbonate involves the extraction of isosorbide from maize starch for
the purpose of polymerisation, thereby eliminating the presence of toxic substances.
Lignocellulosic biomass has attracted considerable attention in the realm of eco-­
friendly materials and 3D printing, primarily owing to its abundant availability and
sustainable nature. Lignocellulosic biomass, which consists of cellulose, hemicel-
lulose, lignin, and various other constituents, has found extensive application in
diverse fields including paper production, biofuel generation, and the development
of biocomposite materials (Yoo et al., 2020). Biomass exhibits biodegradability and
actively engages in the carbon cycle, rendering it more ecologically sustainable in
contrast to materials derived from petroleum. Furthermore, it is frequently observed
that products derived from biomass possess a certain level of biodegradability. The
utilisation of biomass has been expanded due to recent advancements in the under-
standing of its characteristics (Van Wijk & van Wijk, 2015). The utilisation of
biomass-­derived materials in the field of 3D printing presents a promising avenue
for the realisation of a bioeconomy that is both sustainable and renewable.

2 3D Printing Techniques for Biomass-Derived Materials

3D printing techniques have been increasingly explored for the fabrication of


biomass-­derived materials. These techniques allow for the precise and customisable
manufacturing of objects using renewable biomass resources. Here are a few nota-
ble 3D printing techniques employed with biomass-derived materials:
1. Fused Deposition Modelling (FDM): Fused deposition modelling (FDM), also
known as fused filament fabrication (FFF), is a widely used 3D printing technol-
ogy that has revolutionised the world of rapid prototyping and additive manufac-
turing. FDM is known for its simplicity, versatility, and cost-effectiveness,
making it accessible to both professionals and hobbyists. The principle behind
FDM involves the precise deposition of melted thermoplastic materials layer by
layer to build a three-dimensional object. It operates on the concept of an extru-
sion process, where a filament of thermoplastic material is fed into a heated
nozzle. The filament is melted within the nozzle, and the molten material is then
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 155

extruded onto a build platform or the previously deposited layers. The material
quickly solidifies upon deposition, creating a strong bond with the previous lay-
ers. One of the key advantages of FDM is its versatility in material selection.
There is a wide range of thermoplastic filaments available for FDM, each with
its unique properties and characteristics. Common materials used in FDM
include PLA, ABS, PETG, nylon, and TPU. This allows for the production of
objects with various mechanical properties, such as flexibility, strength, and heat
resistance (Tran et al., 2017).
FDM printers are available in a range of sizes and price points, making them
suitable for a wide range of applications, from small-scale hobby projects to
industrial manufacturing. The applications of FDM are extensive and diverse. It
is widely used in product development, enabling rapid prototyping and iterative
design processes. Applications of FDM are found in various industries, includ-
ing automotive, aerospace, architecture, healthcare, education, and consumer
goods. While FDM offers numerous advantages, it is important to note that
printed objects may exhibit certain limitations. These include visible layer lines,
lower resolution compared to other 3D printing technologies, and the need for
support structures for complex geometries. FDM has democratised 3D printing,
providing a practical and accessible method for turning digital designs into tan-
gible objects. With its simplicity, versatility, and affordability, FDM continues to
play a significant role in shaping the future of manufacturing and design (Nguyen
et al., 2018; Ji et al., 2020).
2. Selective Laser Sintering (SLS): Selective laser sintering (SLS) is an additive
manufacturing technique that utilises a high-power laser to selectively fuse pow-
dered materials, typically polymers or metals, to create three-dimensional
objects. SLS is known for its versatility, as it enables the fabrication of complex
geometries and functional prototypes with excellent mechanical properties. The
SLS process begins with a thin layer of powdered material spread evenly across
a build platform. A high-powered laser selectively scans the cross-section of the
object to be printed, selectively sintering or melting the powder particles in the
desired areas. The heat generated by the laser fuses the particles together, form-
ing a solid layer. The build platform then moves down, and a new layer of pow-
der is applied on top. This layer-by-layer process is repeated until the entire
object is formed within the powder bed (Palaganas et al., 2017). One of the sig-
nificant advantages of SLS is its ability to produce parts without the need for
support structures. As the surrounding powder acts as a support, it eliminates the
need for manual removal of supports after printing, reducing post-processing
requirements. Additionally, the powder bed acts as a thermal insulator, allowing
for efficient cooling and minimising warping or distortion in the printed parts.
SLS offers a wide range of material options, including various thermoplastics,
such as nylon, polycarbonate, and polypropylene, as well as metal powders. This
versatility allows for the creation of objects with diverse mechanical properties,
such as strength, flexibility, and heat resistance. Furthermore, SLS enables the
fabrication of composite materials by incorporating additives, such as reinforc-
ing fibres or fillers, into the powder (Li et al., 2019).
156 C. Chauhan et al.

3. Binder Jetting: Binder jetting is an additive manufacturing technique that


involves the precise deposition of a liquid binding agent onto powdered materi-
als, layer by layer, to create three-dimensional objects. It is a versatile and effi-
cient method that has gained significant attention in the field of 3D printing for
its ability to produce complex parts with a wide range of materials. The binder
jetting process begins with a thin layer of powdered material, which can be met-
als, ceramics, or even polymers, spread across a build platform. A print head or
nozzle then selectively deposits a liquid binding agent onto the powdered layer,
binding the particles together in the desired areas. This process is repeated layer
by layer until the entire object is formed. Once the printing is complete, the
object is typically subjected to post-processing steps, such as curing, sintering,
or infiltrating, to enhance its mechanical properties and surface finish (Infanger
et al., 2019).
One of the key advantages of binder jetting is its ability to produce parts at a
relatively high speed, making it suitable for both prototyping and production
applications. It allows for the creation of complex geometries and intricate inter-
nal features that may be challenging to achieve with traditional manufacturing
methods. Additionally, binder jetting offers excellent material efficiency, as the
unbound powder can be reused for future prints, reducing waste and cost. The
versatility of binder jetting is another notable aspect. It can accommodate a wide
range of materials, including metals, ceramics, sand, and composite powders.
This opens up possibilities for diverse applications in industries such as automo-
tive, aerospace, healthcare, and consumer goods. Furthermore, binder jetting
enables the production of large-scale objects, as the layer-by-layer process is not
constrained by the size limitations often encountered in other 3D printing tech-
niques. Additionally, the mechanical properties of binder jetted parts may not be
as robust as those achieved through traditional manufacturing methods. However,
ongoing research and development in materials and process optimisation aim to
overcome these limitations and improve the overall quality of Binder Jetted
objects (Ji et al., 2020).
4. Direct Ink Writing (DIW): Direct ink writing (DIW) is an advanced 3D printing
technique that enables the precise deposition of highly viscous or shear-thinning
materials, known as inks, to fabricate complex three-dimensional structures
(Astm, 2015). DIW offers the ability to print materials with a wide range of vis-
cosities, including pastes, gels, and even suspensions containing functional par-
ticles. Unlike traditional 3D printing methods that utilise solid filaments or
powders, DIW relies on the extrusion of inks through a fine nozzle or syringe
(Markstedt et al., 2017). The inks are typically composed of a base material,
such as polymers, ceramics, metals, or composites, combined with solvents or
binders to achieve the desired flow properties. The ink’s rheological behaviour,
including viscosity and shear-thinning properties, is crucial in DIW to allow for
controlled extrusion, and precise deposition offers several advantages over other
3D printing techniques (Hausmann et al., 2020). Firstly, it enables the fabrica-
tion of structures with complex geometries and intricate internal features, mak-
ing it suitable for applications in tissue engineering, soft robotics, microfluidics,
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 157

and electronics. Secondly, DIW provides the flexibility to print multi-material or


gradient structures by extruding different inks simultaneously or sequentially.
This capability expands the range of functional properties that can be achieved
in the final printed object. Finally, DIW allows for the direct writing of materials
without the need for additional post-processing steps, enhancing the efficiency
of the fabrication process. The applications of DIW are diverse and rapidly
expanding. In the field of bioprinting, DIW enables the precise deposition of
bioinks containing living cells, offering the potential for the fabrication of func-
tional tissues and organs. In the realm of advanced materials, DIW allows for the
creation of complex ceramic or composite structures with tailored properties.
DIW also finds applications in the fabrication of microfluidic devices, sensors,
flexible electronics, and customised products (Pattinson & Hart, 2017).

3 Biomass-Derived 3D Printing Materials

Biomass-derived 3D printing materials refer to raw materials used in 3D printing


that are derived from renewable biomass sources. These materials offer a more sus-
tainable and environmentally friendly alternative to traditional fossil-based materi-
als. Here are some examples of biomass-derived materials used in 3D printing
(Fig. 1).
1. Lignin-Based Materials: Lignin-derived materials have attracted considerable
interest as a promising category of materials for additive manufacturing owing to
their distinctive characteristics and potential for environmentally friendly manu-
facturing processes. Lignin, an intricate naturally occurring polymer obtained
from the cell walls of plants, ranks as the second most prevalent biopolymer
globally (Ragauskas et al., 2014). Consequently, it presents a compelling substi-
tute for materials derived from petroleum. Lignin exhibits a multitude of advan-
tageous attributes that render it suitable for implementation in 3D printing
endeavours. The material demonstrates favourable mechanical strength, elevated
thermal stability, and notable resistance to UV radiation, rendering it appropriate
for a wide range of functional and structural components. Lignin possesses the
additional benefit of being a renewable and biodegradable substance, thereby
aligning with the increasing need for sustainable manufacturing practices.
Different approaches have been investigated in order to employ lignin in the
context of 3D printing. Lignin has the potential to undergo processing to achieve
a desirable form, such as lignin powder or pellets. These forms can subsequently
be combined with other polymers, such as polylactic acid (PLA) or acrylonitrile
butadiene styrene (ABS), in order to create composite filaments suitable for
printing. The inclusion of lignin in the composition of printed objects contributes
to the improvement of their mechanical properties and promotes their environ-
mental sustainability (Vaidya et al., 2019).
158 C. Chauhan et al.

Fig. 1 Classification of Bio-Based Polymers

An alternative method entails the direct application of inks derived from lig-
nin in processes such as inkjet printing or DIW. Lignin-based inks have the
capability to be prepared through the dissolution or dispersion of lignin in
appropriate solvents or binders, thereby facilitating the accurate application and
sequential construction of structures with a high lignin content. This approach
provides the capability to generate intricate geometries and gradients with spe-
cific functionalities. The scope of these items encompasses a wide range of con-
sumer products and packaging, as well as automotive parts, construction
components, and biomedical devices. The inherent UV stability and thermal
resistance of lignin render it well-suited for outdoor applications, while its bio-
degradability enhances its attractiveness for sustainable packaging solutions
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 159

(Ji et al., 2020). The utilisation of lignin for 3D printing presents several chal-
lenges, which encompass the need to optimise the formulation of the material,
ensure its compatibility with printing processes, and achieve a consistent level of
print quality. The intricate composition and diverse characteristics of lignin,
which are influenced by its origin and processing techniques, require meticulous
analysis and adjustment to achieve optimal functionality. However, current
research and development endeavours are investigating innovative approaches
for the extraction, purification, and modification of lignin. Additionally, there is
a focus on the creation of additives and blends derived from lignin, with the aim
of improving printability and material characteristics. With continued progress,
materials derived from lignin possess the capacity to facilitate the sustainable
progression of 3D printing, presenting environmentally conscious alternatives
for various applications and diminishing the dependence on plastics derived
from fossil fuels (Ibrahim et al., 2019).
2. Cellulose-Based Materials: Extensive research and study have been conducted
on cellulose, a naturally occurring polymer that is widely abundant in the cell
walls of plants. Cellulose can manifest in diverse forms, including cellulose
nanofibrils, cellulose nanocrystals, and bacterial cellulose, contingent upon the
isolation methods employed and the sources from which it is derived (Credou &
Berthelot, 2014). Cellulose possesses a hydrophilic character attributed to the
plentiful hydroxyl groups present on its surface. This property enables cellulose
to readily disperse in water, rendering it well-suited for implementation in direct
ink writing (DIW) 3D printing techniques. Consequently, cellulose can be
employed to fabricate hydrogels utilised in diverse applications. The rheological
characteristics of the cellulose-based suspension play a critical role in ensuring
the stability of hydrogel structures during DIW 3D printing. These properties,
such as shear-thinning behaviour, adequate yield stress, finite elastic modulus,
and rapid elastic recovery, are of utmost importance.
Lignocellulosic fibres, renowned for their abundant supply, exceptional spe-
cific mechanical properties, and environmentally friendly nature, have been
extensively employed in the formulation of polymeric-based composites.
Composite materials are frequently formed by combining these fibres with poly-
mers derived from petroleum or bio-based sources. Petroleum-derived polymers
such as epoxy, polyurethane, polypropylene, polyethylene, and acrylonitrile
butadiene styrene (ABS) have demonstrated efficacy in generating composite
materials possessing favourable mechanical and physical characteristics. The
utilisation of biodegradable thermoplastic polymers, including but not limited to
polylactic acid (PLA), polyhydroxyalkanoates (PHA), polyglycolic acid (PGA),
starch, and bio-derived thermoset resins, presents novel opportunities for the
development of truly sustainable materials. The development of lignocellulosic
fibre composites for diverse commercial applications is being propelled by the
potential for recyclability and the integration of bio-derived materials (Aniq
et al., 2020).
Various methods have been utilised in composite manufacturing to incorpo-
rate different types of cellulose, including NFC, CNC, MF, BC, and
160 C. Chauhan et al.

RC. ­Formulations based on polylactic acid have garnered considerable interest


in the pursuit of enhancing their mechanical and physical properties. In the realm
of thermoplastics, polylactic acid has demonstrated superior tensile performance
within cellulose-reinforced composite materials. Nevertheless, there are still cer-
tain obstacles that must be overcome in relation to PLA-based composites,
including issues pertaining to fragility, low impact strength, and thermal stability.
The hydrophilic nature of cellulose presents a significant obstacle in the
development of composites due to its impact on the dispersal and interfacial
adhesion with hydrophobic matrices, such as various biopolymers. Numerous
methodologies have been investigated in order to enhance the compatibility of
cellulose with hydrophobic matrices through physical or chemical surface modi-
fications. The utilisation of nanoscale iterations of cellulose, such as bacterial
cellulose, nano-fibrillated cellulose (NFC), and cellulose nanocrystals (CNC),
has the potential to augment interfacial adhesion through the provision of a
greater surface area relative to volume. Nevertheless, the dispersion of cellulose
within composites is still constrained by chemical incompatibility. Consequently,
chemical alterations, such as acetylation, esterification, and grafting, have been
utilised to augment the compatibility between micro- and nanocellulose fibres
and polymeric matrices (Hausmann et al., 2020).
3. Bio-Based Polymers: A diverse range of bio-based polymers that are obtained
from sustainable biomass sources have the potential to be utilised in the process
of 3D printing. These materials encompass polylactic acid (PLA), which is
derived from corn starch or sugarcane, and polyhydroxyalkanoates (PHA),
which are synthesised by microorganisms using renewable feedstocks. Bio-­
based polymers exhibit a diverse array of characteristics, including but not lim-
ited to biodegradability, flexibility, and strength, rendering them well-suited for
a variety of 3D printing applications. Biopolymers consist of interconnected
monomeric units that are joined through covalent bonds, resulting in the forma-
tion of polymeric biomolecules. In contrast to synthetic polymers, biopolymers
exhibit a sophisticated molecular architecture that results in a precisely delin-
eated three-dimensional configuration. The term “bio” in reference to biopoly-
mers denotes their intrinsic capacity to undergo natural degradation via biological
mechanisms. Biopolymers play a significant role in serving as biodegradable
materials, thereby presenting an attractive substitute for polymers derived from
fossil fuels. This characteristic of biopolymers is particularly noteworthy as it
effectively addresses pressing environmental issues, contributing to the conser-
vation of our planet. In contrast, synthetic polymers are artificial substances that
are not capable of undergoing biodegradation and are produced by humans
through the utilisation of petrochemicals. The lifecycle of these materials fol-
lows a linear trajectory, commencing with the extraction and utilisation of finite
oil resources. These resources are subsequently transformed into polymers,
which are then employed in the manufacturing of various products (Phiri et al.,
2023). Ultimately, these products are disposed of at the end of their useful life.
The utilisation of a linear approach in various industries has been found to have
a substantial impact on global pollution levels, while also being heavily reliant
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 161

on finite fossil fuel resources. In contrast, biopolymers are in alignment with a


circular paradigm in terms of material design, manufacturing, application, and
disposal. The materials are integrated into a more sustainable cycle, taking into
account the complete life cycle of the material. The adoption of a circular
approach serves to decrease dependence on fossil fuels, enhance resource effi-
ciency, and mitigate environmental consequences (Visco et al., 2022). The adop-
tion of biopolymers presents an opportunity to transition towards a paradigm
that is characterised by enhanced sustainability and greater environmental
friendliness.

3.1 Polylactic Acid

Polylactic acid (PLA) is a biodegradable polymer that is derived from renewable


sources such as maize, sugar feedstock, maize, wheat, and agricultural waste. It is
produced through a process involving fermentation and controlled polymerisation
of lactic acid monomers. The manufacturing process of PLA necessitates a lower
energy input and results in reduced greenhouse gas emissions in comparison with
various synthetic polymers (Sabee et al., 2021). This characteristic positions PLA as
one of the limited biopolymers that are manufactured on a significant scale, enabling
it to effectively rival synthetic polymers and effectively supplant them in diverse
applications. PLA is classified as a thermoplastic material owing to its response to
changes in temperature. Thermoplastic materials, such as PLA, possess the ability
to undergo repeated cycles of heating to their respective melting points, subsequent
cooling, and subsequent reheating, while maintaining their structural integrity with-
out experiencing substantial degradation. The aforementioned attribute possesses
significant value due to its facilitation of efficient injection moulding processes and
subsequent recyclability. During the process of crystallisation, PLA undergoes the
formation of three distinct structural forms, namely, α, β, and γ (Zwawi, 2021). The
α-form is generated via the process of cold crystallisation, whereas the β-form is a
consequence of mechanical effects and deformation. The high crystallisation rate of
PLA can be attributed to the rapid expansion of semi-crystalline domains. The prop-
erties of a material, such as melting point, tensile strength, impact strength, hard-
ness, and stiffness, are influenced by the degree of crystallinity. PLA is widely
recognised for its advantageous characteristics, such as its biocompatibility, moder-
ate mechanical strength, and affordability (Sreekumar et al., 2021). Nevertheless,
PLA films exhibit characteristics such as brittleness, fragility, and low toughness.
The aforementioned attributes impose constraints on their appropriateness for high-­
stress scenarios and present difficulties in implementing surface modifications. PLA
encounters difficulties related to thermal instability, wherein it undergoes molecular
weight reduction and degradation of ester linkages when subjected to elevated tem-
peratures. Even temperatures that are lower than the melting point have the potential
to expedite this process of degradation. Various techniques, including esterification
reactions, hydrolysis, depolymerisation, and oxidative degradation, are utilised in
162 C. Chauhan et al.

order to address the issue of thermal degradation. The degree of degradation is con-
tingent upon various factors, including the level of moisture, concentration of lactic
acid, and size of particles (Phiri et al., 2023).

3.2 Polyhydroxyalkanoates (PHAs)

Polyhydroxyalkanoates (PHAs) represent a category of biodegradable polyesters


that are naturally synthesised via bacterial fermentation. Bacterial organisms
employ the synthesis of PHAs as a means of survival, achieved through the accumu-
lation of PHA granules during the process of fermentation. Short-chain PHAs are
composed of three to five monomers, specifically polyhydroxybutyrate (PHB),
polyhydroxyvalerate (PHV), and their copolymers. PHB, renowned for its excep-
tional mechanical performance capabilities, has been extensively investigated and
widely employed within this particular subgroup of polymers. Medium-chain-­
length polyhydroxyalkanoates (mcl-PHAs) are synthesised by employing mono-
mers that consist of carbon atoms ranging from 6 to 14. Examples of such monomers
include polyhydroxyoctanoate, polyhydroxylhexanoate, and polyhydroxydecanoate
(Miu et al., 2022).
The mechanical and biological compatibility of PHAs can be altered through
various methods such as blending, surface modification, or incorporation with other
polymers, enzymes, or inorganic materials. This broadens the scope of their poten-
tial applications. In addition, the characteristics of PHAs can be improved by incor-
porating other materials through physical blending with nanoparticles and
nanofillers, resulting in the formation of multiphase materials, as well as undergo-
ing chemical modifications. The primary objective of endeavours focused on
enhancing and adapting PHAs is to enhance their competitiveness relative to con-
ventional plastics in terms of physicochemical characteristics and manufacturing
expenses. Cataldi et al. (2020) conducted an independent investigation wherein they
employed a melt blending technique to combine graphene nanoplatelets with PHA
biopolymer. This amalgamation was found to significantly improve the thermal sta-
bility, thermal conductivity, and electromagnetic shielding capabilities of the result-
ing composite material. Consequently, this enhanced composite exhibits promising
potential for utilisation in electric applications.

4 Polysaccharide-Based Materials

4.1 Chitosan

Chitosan, a natural polymer derived from shellfish, is obtained through the alkaline
N-deacetylation process of chitin. This versatile material finds extensive use in the
realm of 3D printing. The efficacy of chitosan is contingent upon its acetylation
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 163

degree and molecular weight, both of which exert influence over a range of charac-
teristics including printability, biodegradability, shape fidelity, and mechanical
strength. Nevertheless, the utilisation of pure chitosan constructs is hindered by the
difficulties encountered in attaining accurate geometries and structural stability
(Shahbazi et al., 2016; Hospodiuk et al., 2017). These challenges arise from the
hydrophilic properties of chitosan and the tendency of three-dimensional structures
to undergo shrinkage, thereby impeding the fabrication of printed architectures. In
order to address these constraints, considerable endeavours have been undertaken to
formulate printable inks based on chitosan for diverse applications. These endeav-
ours encompass the amalgamation of chitosan with various other substances, includ-
ing PLA, silk protein, gelatine, CMC, and gellan (Robinson et al., 2020). Wu et al.
(2017) devised a three-dimensional (3D) printing methodology to fabricate hydro-
gel scaffolds utilising chitosan as the primary material, featuring meticulously
arranged microfiber architectures. The scaffolds that were produced demonstrated a
failure strain of approximately 400% and a maximum strength of around 7.5 MPa.
In a separate investigation conducted by Zhang et al., the authors employed physi-
cally ground silk powders as a means to enhance the mechanical properties of chi-
tosan hydrogels. This intervention led to notable enhancements in compressive
modulus, printing precision, and construct stability when compared to hydrogels
composed solely of chitosan.
The researchers were able to successfully produce scaffolds with intricate geom-
etries and feature sizes of approximately 50 μm. In a similar vein, Zolfagharian
et al. (2018) focused on optimising printing adjustments and variables to develop a
polyelectrolyte complex hydrogel utilising chitosan and gelatine. The resulting
hydrogel exhibited superior maximum deflection and improved properties when
compared to cast films. Chitosan colloidal dispersion has been employed as a bioink
in the context of 3D bioprinting as well. Recent research has examined different
processing parameters associated with bioinks based on chitosan, encompassing
aspects such as flow characteristics, structural integrity, compatibility with living
organisms, and the rate of solvent evaporation. A separate investigation was con-
ducted wherein a scaffold resembling the shape of an ear was produced through the
employment of a combination of chitosan and PEGDA via the technique of stereo-
lithography. The enhancement of mechanical strength, printing performance, and
cell adhesion was achieved through the manipulation of chitosan’s molecular weight
and the proportion of chitosan, PEGDA, and photosensitiser. The incorporation of
chitosan into PEGDA resulted in an enhancement of the flow characteristics of the
system, thereby allowing for a reduction in the required concentration of PEGDA
(from 30% to 6.5% w/v) during the printing procedure. Furthermore, the manipula-
tion of the feed ratios between chitosan and PEGDA had a significant effect on the
swelling behaviour and elastic modulus of the fabricated scaffolds (Morris
et al., 2017).
164 C. Chauhan et al.

4.2 Starch

Starch, a natural polymer known for its favourable attributes and cost-effectiveness,
finds extensive application within the pharmaceutical, bioengineering, and food
sectors. The substance is derived from a variety of sources including cassava, corn,
potato, rice (Chen et al., 2019), and wheat and has been utilised in the field of 3D
printing to improve flow characteristics and extrusion capabilities (Yang et al.,
2018a, b). Starch, regardless of its state as granular or gelatinised, can be incorpo-
rated into biopolymer systems to generate printed structures that exhibit enhanced
printability, biodegradability, precise geometries, and desired structural attributes.
In order to attain prints of superior resolution and enhanced shape retention, starch
has been effectively blended with various biopolymers, such as cellulose, oat and
faba bean proteins, and carrageenan xanthan. These biopolymers possess favourable
processability characteristics and provide sufficient structural integrity. The investi-
gation of the correlation between the flow characteristics of starches produced
through 3D printing and their subsequent printing efficacy has been undertaken by
researchers. Promising outcomes have been observed in the 3D printing of struc-
tures using concentrated starches that demonstrate shear-thinning characteristics
and strain responsiveness. As an example, rice starch with a weight percentage of
15–25% that has undergone gelatinisation at a temperature of 80 °C, corn starch
with a weight percentage of 20–25% processed at a temperature of 75 °C, and potato
starch with a weight percentage of 15–20% treated at a temperature of 70 °C have
exhibited suitable values for flow stress, yield stress, and elastic modulus, respec-
tively. Moreover, the incorporation of potato starch into lemon juice hydrogels has
exerted an impact on the rheological characteristics and mechanical attributes,
underscoring the necessity of adjusting the nozzle height relative to the nozzle
diameter (Yang et al., 2018a, b).

5 The Utilisation of 3D Printing Technology in Conjunction


with Biomass Materials

5.1 Biomedical

Biomass materials possess notable advantages within the biomedical domain owing
to their inherent biocompatibility, biodegradability, and minimal toxicity. Biomass
3D printing is utilised in diverse biomedical domains, encompassing the production
of synthetic scaffolds (Yang et al., 2018a, b), simulated tissues (Athukoralalage
et al., 2019), the promotion of wound healing (Kanikireddy et al., 2020), and the
development of drug delivery systems (Surini et al., 2023; Zamboulis et al., 2022).
The utilisation of novel soybean oil epoxy acrylates in biomedical scaffolds was
investigated by Miao et al. (2016). The material demonstrated shape memory prop-
erties and successfully underwent cytotoxicity assessments. The study exhibited
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 165

enhanced adhesion of human bone marrow mesenchymal stem cells (hMSCs) and
additional advantageous characteristics in comparison with conventional polyethyl-
ene glycol diacrylate (PEGDA). This study made a significant contribution to the
investigation of the integration of vegetable oils and other sustainable chemicals in
order to facilitate the implementation of 4D printing. The utilisation of biomass
materials in conjunction with biomass-derived polylactic acid (PLA) yields a com-
posite possessing attributes that are both non-toxic and environmentally sustainable.
In their study, Calì et al. (2020) employed polylactic acid as the substrate material
and integrated antibacterial hemp reinforcement to fabricate materials with proper-
ties resembling silk.
Nanocellulose has the potential to be utilised in the fabrication of transparent
films that offer a moist environment conducive to wound healing, as well as elastic
gels exhibiting bio-responsive characteristics. In their study, Rees et al. (2015)
employed carboxymethylated periodate-oxidised nanocellulose as a material for the
production of three-dimensional printed wound dressings. These dressings were
specifically designed to inhibit bacterial growth, rendering them appropriate for use
in medical contexts. The researchers such as Leppiniemi et al. (2017) successfully
fabricated wound dressings using a combination of nanocellulose and alginate
hydrogel, which were found to be compatible with 3D printing technology. In their
study, Xu et al. (2019) developed biomimetic inks through the combination of cel-
lulose nanofibers (CNFs) and cross-linkable hemicellulose derivatives. These inks
were employed as a substrate for the cultivation of human skin cells and pancreatic
tumour cell cultures. The scaffolds that were obtained demonstrated favourable bio-
compatibility and maintained their original shape accurately. Avila and colleagues
(2016) fabricated auricular constructs by incorporating human nasal cartilage (hNC)
into a composite material composed of nano-fibrillated cellulose and alginate (NFC-­
A). The integration of non-toxic and biocompatible biomass materials with tailored
3D printing techniques represents a novel methodology within the realm of medical
applications.

5.2 Electronics

In contrast to traditional processing methods, the utilisation of 3D printing technol-


ogy provides the capability to manufacture highly complex shapes. Biomass materi-
als have been usually utilised in the electronics industry, facilitating the
manufacturing of diverse products including supercapacitors (Zhang et al., 2018),
wearable jewellery (Zhu et al., 2022), and sensors (Hamad, 2016). Nanocellulose,
renowned for its inherent ability to self-assemble and achieve a high degree of struc-
tural alignment, demonstrates enhanced thermal conductivity and is utilised in vari-
ous fields such as supercapacitors, lithium-ion batteries, and the thermal regulation
of solar cells. The acquisition of composites with elevated thermal conductivity can
be achieved by the amalgamation of cellulose nanofibers (CNF) with carbon nano-
tubes, graphene, and inorganic nitrides (Ee & Li, 2021). Cao et al. (2019) employed
166 C. Chauhan et al.

carbon nanofiber (CNF) composites as the primary constituents for the fabrication
of three-dimensional printed lithium metal batteries with enhanced performance. In
their study, Wang et al. (2022) utilised the selective laser sintering printing method
to fabricate porous wood precursors derived from biomass. These precursors were
subsequently combined with carbon thermal reduction techniques to achieve the
integration of structure and functionality for the purpose of electromagnetic wave
absorption. The materials commonly employed as substrates for flexible electronics
are typically paper or polymer films. In the context of sustainable development, it is
imperative to transition raw materials used in electronic energy storage towards
renewable and ecologically sound alternatives. Lignin and cellulose ether groups
possess the potential to function as viable feedstocks for the fabrication of elec-
tronic energy storage components using 3D printing technology. In their study,
Huang et al. (2019) fabricated hydrogels that exhibited remarkable electrical con-
ductivity and thermal sensitivity. This was achieved through the utilisation of
hydroxyethyl cellulose (HEC) and polyvinyl alcohol (PVA) as the primary constitu-
ents, borax as the cross-linking agent, and lignin as the plasticising agent. The
stretchable electrodes and diaphragms for lithium-ion batteries were developed by
Li’s research group at the University of Maryland, USA (Qian et al., 2022). This
was achieved through the utilisation of nanocellulose, carbon nanotubes, and active
materials combined in aqueous inks. The printed serpentine structures displayed a
notable degree of deformability. The robust nanoscale structure formed by the
strong interaction between cellulose and nanocellulose, combined with their high
aspect ratio, exhibits considerable promise for applications in wearable devices and
electronic energy storage.

5.3 Construction Field

The utilisation of 3D printing technology in the construction industry presents


numerous benefits, including but not limited to customisation, effective material
usage, and autonomous shaping, particularly when applied to the printing of large-­
scale components. This attribute renders it a valuable instrument for preliminary
design models, thereby affording designers with adaptability in their creative meth-
odology. Furthermore, the utilisation of 3D printing technology extends to the con-
struction industry, where it can be employed for the creation of building foundations
and the fabrication of various building materials. The method proposed by Yoshida
et al. (2015) involved the utilisation of 3D printing technology to fabricate a build-
ing model, resulting in the successful production of a venue model. The construc-
tion materials employed in their methodology comprised ulna, a variety of timber,
and adhesive derived from Chinese fir. The successful 3D printing of a building base
was accomplished by Zhao et al. (2019) through the utilisation of poplar fibre and
PLA as raw materials. The characteristics of the printed substrate exhibited stability
and were subject to control. The utilisation of concrete in the field of construction
has resulted in a reduction in the utilisation of soil-based materials and a concurrent
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 167

rise in the emission of greenhouse gases. In their research investigation, Alqenaee


and Memari (2022) employed a composite mixture of straw, water, sand, and clay
with the aim of examining environmentally friendly alternatives within the domain
of construction materials. Through the manipulation of material ratios, the research-
ers successfully improved the tensile and compressive characteristics of the printed
mixture, indicating that the altered material exhibits promise for application in the
construction of printed houses.

6 Conclusions and Perspectives

In contemporary times, there have been notable progressions in the field of 3D


printing technologies, which have garnered attention as a potentially advantageous
approach to manufacturing. These advancements have demonstrated several bene-
fits, including the mitigation of waste production, decreased energy usage, and the
capacity to construct intricate and intricate structures. In response to the difficulties
presented by petroleum-derived materials, there has been a growing body of research
focused on the utilisation of lignocellulosic biomass within the realm of 3D print-
ing. This approach effectively addresses challenges associated with limited resources
and environmental consequences. In the present scenario, various substances includ-
ing cellulose, lignin, starch, and whole biomass have been employed in diverse 3D
printing methodologies, such as fused deposition modelling, stereolithography,
binder jetting, and direct ink writing. These technologies facilitate the utilisation of
economically viable biomass-derived materials for the production of composites
that are suitable for various applications in industries such as construction, biomedi-
cal, pharmaceutical, and food. The comprehension of the arrangement of biomass
constituents and the application of supramolecular structures, including crystallin-
ity, material anisotropy, and interfacial interactions, are crucial factors to be taken
into account in this burgeoning area of study. A thorough investigation is required to
understand the hierarchical formation of the final product. In addition, the optimisa-
tion of processing parameters, including printing resolution and production rate,
plays a critical role in enhancing the competitiveness of 3D printing in comparison
with conventional fabrication methods. The investigation of in situ characterisation
techniques, akin to those employed in metal 3D printing, presents an opportunity to
augment the printing precision within the realm of biomass-­derived materials.

References

Alqenaee, A., & Memari, A. (2022). Experimental study of 3D printable cob mixtures.
Construction and Building Materials, 324, 126574.
Astm, I. (2015). ASTM52900-15 standard terminology for additive manufacturing—General prin-
ciples—Terminology. ASTM International, West Conshohocken, PA, 3(4), 5.
168 C. Chauhan et al.

Athukoralalage, S. S., Balu, R., Dutta, N. K., & Roy Choudhury, N. (2019). 3D bioprinted
nanocellulose-based hydrogels for tissue engineering applications: A brief review. Polymers,
11(5), 898.
Ávila, H. M., Schwarz, S., Rotter, N., & Gatenholm, P. (2016). 3D bioprinting of human
chondrocyte-­laden nanocellulose hydrogels for patient-specific auricular cartilage regenera-
tion. Bioprinting, 1, 22–35.
Calì, M., Pascoletti, G., Gaeta, M., Milazzo, G., & Ambu, R. (2020). New filaments with natural
fillers for FDM 3D printing and their applications in biomedical field. Procedia Manufacturing,
51, 698–703.
Cao, D., Xing, Y., Tantratian, K., Wang, X., Ma, Y., Mukhopadhyay, A., Cheng, Z., Zhang, Q., Jiao,
Y., Chen, L., & Zhu, H. (2019). 3D printed high-performance lithium metal microbatteries
enabled by nanocellulose. Advanced Materials, 31(14), 1807313.
Cataldi, P., Steiner, P., Raine, T., Lin, K., Kocabas, C., Young, R. J., Bissett, M., Kinloch, I. A., &
Papageorgiou, D. G. (2020). Multifunctional biocomposites based on polyhydroxyalkanoate
and graphene/carbon nanofiber hybrids for electrical and thermal applications. ACS Applied
Polymer Materials, 2(8), 3525–3534.
Chen, H., Xie, F., Chen, L., & Zheng, B. (2019). Effect of rheological properties of potato, rice
and corn starches on their hot-extrusion 3D printing behaviors. Journal of Food Engineering,
244, 150–158.
Credou, J., & Berthelot, T. (2014). Cellulose: From biocompatible to bioactive material. Journal of
Materials Chemistry B, 2(30), 4767–4788.
Ee, L. Y., & Li, S. F. Y. (2021). Recent advances in 3D printing of nanocellulose: Structure, prepa-
ration, and application prospects. Nanoscale Advances, 3(5), 1167–1208.
Hamad, W. Y. (2016). Photonic and semiconductor materials based on cellulose nanocrys-
tals. In Cellulose chemistry and properties: Fibers, nanocelluloses and advanced materials
(pp. 287–328). Springer.
Hausmann, M. K., Siqueira, G., Libanori, R., Kokkinis, D., Neels, A., Zimmermann, T., & Studart,
A. R. (2020). Complex-shaped cellulose composites made by wet densification of 3D printed
scaffolds. Advanced Functional Materials, 30(4), 1904127.
Hospodiuk, M., Dey, M., Sosnoski, D., & Ozbolat, I. T. (2017). The bioink: A comprehensive
review on bioprintable materials. Biotechnology Advances, 35(2), 217–239.
Huang, S., Shuyi, S., Gan, H., Linjun, W., Lin, C., Danyuan, X., Zhou, H., Lin, X., & Qin, Y. (2019).
Facile fabrication and characterization of highly stretchable lignin-based hydroxyethyl cellu-
lose self-healing hydrogel. Carbohydrate Polymers, 223, 115080.
Ibrahim, F., Mohan, D., Sajab, M. S., Bakarudin, S. B., & Kaco, H. (2019). Evaluation of the
compatibility of organosolv lignin-graphene nanoplatelets with photo-curable polyurethane in
stereolithography 3D printing. Polymers, 11(10), 1544.
Infanger, S., Haemmerli, A., Iliev, S., Baier, A., Stoyanov, E., & Quodbach, J. (2019). Powder
bed 3D-printing of highly loaded drug delivery devices with hydroxypropyl cellulose as solid
binder. International Journal of Pharmaceutics, 555, 198–206.
Ji, A., Zhang, S., Bhagia, S., Yoo, C. G., & Ragauskas, A. J. (2020). 3D printing of biomass-­
derived composites: Application and characterization approaches. RSC Advances, 10(37),
21698–21723.
Kanikireddy, V., Varaprasad, K., Jayaramudu, T., Karthikeyan, C., & Sadiku, R. (2020).
Carboxymethyl cellulose-based materials for infection control and wound healing: A review.
International Journal of Biological Macromolecules, 164, 963–975.
Leppiniemi, J., Lahtinen, P., Paajanen, A., Mahlberg, R., Metsä-Kortelainen, S., Pinomaa, T.,
Pajari, H., Vikholm-Lundin, I., Pursula, P., & Hytönen, V. P. (2017). 3D-printable bioactivated
nanocellulose–alginate hydrogels. ACS Applied Materials & Interfaces, 9(26), 21959–21970.
Li, V. C. F., Kuang, X., Mulyadi, A., Hamel, C. M., Deng, Y., & Qi, H. J. (2019). 3D printed cel-
lulose nanocrystal composites through digital light processing. Cellulose, 26, 3973–3985.
Ligon, S. C., Liska, R., Stampfl, J., Gurr, M., & Mülhaupt, R. (2017). Polymers for 3D printing and
customized additive manufacturing. Chemical Reviews, 117(15), 10212–10290.
Utilising Biomass-Derived Composites in 3D Printing to Develop Eco-Friendly… 169

Markstedt, K., Escalante, A., Toriz, G., & Gatenholm, P. (2017). Biomimetic inks based on cellu-
lose nanofibrils and cross-linkable xylans for 3D printing. ACS Applied Materials & Interfaces,
9(46), 40878–40886.
Miao, S., Zhu, W., Castro, N. J., Nowicki, M., Zhou, X., Cui, H., Fisher, J. P., & Zhang, L. G. (2016).
4D printing smart biomedical scaffolds with novel soybean oil epoxidized acrylate. Scientific
Reports, 6(1), 27226.
Miu, D. M., Eremia, M. C., & Moscovici, M. (2022). Polyhydroxyalkanoates (PHAs) as bioma-
terials in tissue engineering: Production, isolation, characterization. Materials, 15(4), 1410.
Morris, V. B., Nimbalkar, S., Younesi, M., McClellan, P., & Akkus, O. (2017). Mechanical prop-
erties, cytocompatibility and manufacturability of chitosan: PEGDA hybrid-gel scaffolds by
stereolithography. Annals of Biomedical Engineering, 45, 286–296.
Nguyen, N. A., Barnes, S. H., Bowland, C. C., Meek, K. M., Littrell, K. C., Keum, J. K., & Naskar,
A. K. (2018). A path for lignin valorization via additive manufacturing of high-performance
sustainable composites with enhanced 3D printability. Science Advances, 4(12), eaat4967.
Palaganas, N. B., Mangadlao, J. D., de Leon, A. C. C., Palaganas, J. O., Pangilinan, K. D., Lee,
Y. J., & Advincula, R. C. (2017). 3D printing of photocurable cellulose nanocrystal compos-
ite for fabrication of complex architectures via stereolithography. ACS Applied Materials &
Interfaces, 9(39), 34314–34324.
Pattinson, S. W., & Hart, A. J. (2017). Additive manufacturing of cellulosic materials with robust
mechanics and antimicrobial functionality. Advanced Materials Technologies, 2(4), 1600084.
Phiri, R., Sanjay, M. R., Siengchin, S., Oladijo, O. P., & Dhakal, H. N. (2023). Development of
sustainable biopolymer-based composites for lightweight applications from agricultural waste
biomass: A review. Advanced Industrial and Engineering Polymer Research, 6(5), 436–450.
Qian, J., Chen, Q., Hong, M., Xie, W., Jing, S., Bao, Y., Chen, G., Pang, Z., Hu, L., & Li, T. (2022).
Toward stretchable batteries: 3D-printed deformable electrodes and separator enabled by nano-
cellulose. Materials Today, 54, 18–26.
Ragauskas, A. J., Beckham, G. T., Biddy, M. J., Chandra, R., Chen, F., Davis, M. F., Davison,
B. H., Dixon, R. A., Gilna, P., Keller, M., & Langan, P. (2014). Lignin valorization: Improving
lignin processing in the biorefinery. Science, 344(6185), 1246843.
Rees, A., Powell, L. C., Chinga-Carrasco, G., Gethin, D. T., Syverud, K., Hill, K. E., & Thomas,
D. W. (2015). 3D bioprinting of carboxymethylated-periodate oxidized nanocellulose con-
structs for wound dressing applications. BioMed Research International, 2015, 925757.
Robinson, T. M., Talebian, S., Foroughi, J., Yue, Z., Fay, C. D., & Wallace, G. G. (2020). Fabrication
of aligned biomimetic gellan gum-chitosan microstructures through 3D printed microflu-
idic channels and multiple in situ cross-linking mechanisms. ACS Biomaterials Science &
Engineering, 6(6), 3638–3648.
Sabee, M. M. S. M., Uyen, N. T. T., Ahmad, N., & Hamid, Z. A. A. (2021). Plastics packaging for
pharmaceutical products. In Reference module in materials science and materials engineering.
Elsevier.
Shahbazi, M., Ettelaie, R., & Rajabzadeh, G. (2016). Physico-mechanical analysis data in support
of compatibility of chitosan/κ-carrageenan polyelectrolyte films achieved by ascorbic acid, and
the thermal degradation theory of κ-carrageenan influencing the properties of its blends. Data
in Brief, 9, 648–660.
Sreekumar, K., Bindhu, B., & Veluraja, K. (2021). Perspectives of polylactic acid from structure to
applications. Polymers from Renewable Resources, 12(1–2), 60–74.
Surini, S., Bimawanti, Y., & Kurniawan, A. (2023). The application of polymers in fabricating 3D
printing tablets by fused deposition modeling (FDM) and the impact on drug release profile.
Pharmaceutical Sciences, 29(2), 156–164.
Tran, P., Ngo, T. D., Ghazlan, A., & Hui, D. (2017). Bimaterial 3D printing and numerical analysis
of bio-inspired composite structures under in-plane and transverse loadings. Composites Part
B: Engineering, 108, 210–223.
170 C. Chauhan et al.

Vaidya, A. A., Collet, C., Gaugler, M., & Lloyd-Jones, G. (2019). Integrating softwood biorefinery
lignin into polyhydroxybutyrate composites and application in 3D printing. Materials Today
Communications, 19, 286–296.
Van Wijk, A. J. M., & van Wijk, I. (2015). 3D printing with biomaterials: Towards a sustainable
and circular economy. IOS Press.
Visco, A., Scolaro, C., Facchin, M., Brahimi, S., Belhamdi, H., Gatto, V., & Beghetto, V. (2022).
Agri-food wastes for bioplastics: European prospective on possible applications in their second
life for a circular economy. Polymers, 14(13), 2752.
Wang, X., Jiang, M., Zhou, Z., Gou, J., & Hui, D. (2017). 3D printing of polymer matrix compos-
ites: A review and prospective. Composites Part B: Engineering, 110, 442–458.
Wang, C., Wu, S., Li, Z., Chen, S., Chen, A., Yan, C., Shi, Y., Zhang, H., & Fan, P. (2022). 3D
printed porous biomass–derived SiCnw/SiC composite for structure–function integrated elec-
tromagnetic absorption. Virtual and Physical Prototyping, 17(3), 718–733.
Wu, Q., Maire, M., Lerouge, S., Therriault, D., & Heuzey, M. C. (2017). 3D printing of micro-
structured and stretchable chitosan hydrogel for guided cell growth. Advanced Biosystems,
1(6), 1700058.
Xu, W., Zhang, X., Yang, P., Långvik, O., Wang, X., Zhang, Y., Cheng, F., Österberg, M., Willför,
S., & Xu, C. (2019). Surface engineered biomimetic inks based on UV cross-linkable wood
biopolymers for 3D printing. ACS Applied Materials & Interfaces, 11(13), 12389–12400.
Yang, E., Miao, S., Zhong, J., Zhang, Z., Mills, D. K., & Zhang, L. G. (2018a). Bio-based poly-
mers for 3D printing of bioscaffolds. Polymer Reviews, 58(4), 668–687.
Yang, F., Zhang, M., Bhandari, B., & Liu, Y. (2018b). Investigation on lemon juice gel as food
material for 3D printing and optimization of printing parameters. Lwt, 87, 67–76.
Yoo, C. G., Meng, X., Pu, Y., & Ragauskas, A. J. (2020). The critical role of lignin in ligno-
cellulosic biomass conversion and recent pretreatment strategies: A comprehensive review.
Bioresource Technology, 301, 122784.
Yoshida, H., Igarashi, T., Obuchi, Y., Takami, Y., Sato, J., Araki, M., Miki, M., Nagata, K., Sakai,
K., & Igarashi, S. (2015). Architecture-scale human-assisted additive manufacturing. ACM
Transactions on Graphics (TOG), 34(4), 1–8.
Zamboulis, A., Michailidou, G., Koumentakou, I., & Bikiaris, D. N. (2022). Polysaccharide 3D
printing for drug delivery applications. Pharmaceutics, 14(1), 145.
Zhang, T., Li, X., Asher, E., Deng, S., Sun, X., & Yang, J. (2018). Paper with power: Engraving
2D materials on 3D structures for printed, high-performance, binder-free, and all-solid-state
supercapacitors. Advanced Functional Materials, 28(37), 1803600.
Zhao, X., Tekinalp, H., Meng, X., Ker, D., Benson, B., Pu, Y., Ragauskas, A. J., Wang, Y., Li, K.,
Webb, E., & Gardner, D. J. (2019). Poplar as biofiber reinforcement in composites for large-­
scale 3D printing. ACS Applied Bio Materials, 2(10), 4557–4570.
Zhu, Y., Qin, J., Shi, G., Sun, C., Ingram, M., Qian, S., Lu, J., Zhang, S., & Zhong, Y. L. (2022).
A focus review on 3D printing of wearable energy storage devices. Carbon Energy, 4(6),
1242–1261.
Zolfagharian, A., Kaynak, A., Khoo, S. Y., & Kouzani, A. Z. (2018). Polyelectrolyte soft actuators:
3D printed chitosan and cast gelatin. 3D Printing and Additive Manufacturing, 5(2), 138–150.
Zwawi, M. (2021). A review on natural fiber bio-composites, surface modifications and applica-
tions. Molecules, 26(2), 404.
Bioenergy Production Using Biomass
Wastes: Challenges of Circular Economy

Vijaya Ilango

1 Introduction

Biomass materials are considered to be the primary carbon sources for the manufac-
ture of a wide range of platform compounds, bioproducts, chemicals, and fuels.
These materials are difficult to break down into minor components and/or change to
an extensive variety of value-added products because of their complicated, diverse,
and stiff structures. It takes extra biomass feedstocks to produce the same quantity
of energy as a standard hydrocarbon fuel since biomass has a low energy density. It
has become increasingly challenging in the twenty-first century to provide clean,
affordable, and consistent energy sources, which are vital from mutually a socioeco-
nomic and environmental viewpoint.
To address these critical issues, biomass is currently the most advantageous
renewable source. Biomass has gained prominence in recent years as the only con-
tinuous carbon source accessible on earth. Biomass is a popular heating and cook-
ing fuel in many nations, particularly in developing countries (de Carraro et al.,
2020). Several developed nations are increasing their use of biomass fuels for power
generation and conveyance since biomass has a low energy density, thereby carbon
dioxide emissions from fossil fuel usage can be minimized. Biomass accounts for
around 5 quadrillion British thermal units (Btu) and approximately 5% of total pri-
mary energy consumption in the United States in 2021 (Adekoya et al., 2023).
Globally, fast economic expansion, increased industrialization, and urbanization are
causing a rise in energy use and environmental degradation (Mathimani & Mallick,
2019). Fuel demand for power, heating, and transportation is now met by finite fos-
sil fuels derived from the remnants of living things over millions of years (Chaudry
et al., 2015).

V. Ilango (*)
Department of Chemistry, Birla Institute of Technology & Science, Pilani, Dubai Campus,
Dubai, UAE

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 171
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_9
172 V. Ilango

• Introduction to Biomass waste


and Bioenergy

Contents • Bioenergy production methods


• Challenges and alternatives of
bioenergy production

Fig. 1 Layout depicting the plan of this chapter

Biomass leftovers and garbage are formed as by-products when the essential raw
materials are cultivated, treated, and expended, as opposed to biomass developed
specifically for energy (Speight & Singh, 2014). Although increased industrializa-
tion has provided numerous benefits to our communities, currently energy exhaus-
tion and waste accumulation have unavoidably become significant problems. At
present, among the scientific community’s most important concerns are biomass
wastes from various sources, such as home leftovers, forestry, and agricultural and
industrial wastes (Nair et al., 2022). Biomass has gained prominence in recent years
as the only continuous carbon source accessible on earth. As a result, it is viewed as
an ideal alternative to fossil fuels. The usage of biomass is a modern technique for
power generation and is regarded as a great neutral resource that reduces CO2 emis-
sions. Biomass is an appealing feedstock for a variety of reasons.
This chapter consists of three parts. In the first part, the introduction to biomass
waste and bioenergy along with their importance will be discussed followed by the
second part that gives an overview of the methods adopted for bioenergy produc-
tion. In the third part of discussion, the challenges faced and alternatives for bioen-
ergy production will be deliberated and emphasized from various research clusters.
The plan of this chapter is given in Fig. 1.

2 Biomass and Bioenergy

Renewable energy plays a significant role in the present and future eras in overcom-
ing and replacing rapidly declining fossil fuel supplies, reducing environmental
harm by regulating greenhouse gas emissions, and controlling pollution-related
environmental challenges. Renewable energy and sustainable development have an
essential link in this regard (Boro et al., 2022). Unlike other ecologically friendly
energy sources such as solar, geothermal, wind, hydropower, and marine, biomass
may directly produce fuel and chemicals.
Renewable energy has the ability to fulfill demand while leaving a significantly
smaller environmental impact, and it may also aid in the resolution of other crucial
issues such as energy security by prompting a dispersed and diverse energy
Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 173

Fig. 2 US major
renewable energy
consumption by sources as Solar, 12.3
of 2021
Wind, 27.4

Biomass, 39.8
Hydroelectric,
18.8

Geothermal
1.7

infrastructure (Kumar, 2020). Biofuels have the capability to lessen transportation-­


related energy and greenhouse gas emissions, but they can also have significant
socioeconomic and environmental implications. A significant increase in farming
and irrigation water usage may be essential, depending on technology, demand, and
crop-growing conditions. In the United States, the major renewable energy con-
sumption by sources as of 2021 is given in Fig. 2. (Center for Sustainable Systems,
2021). Among the total energy consumption of 97.2 quads, the renewables are
found to be 12.2 quads. Quad = quadrillion (1015) Btu. One quad is equivalent to
the annual energy consumption of ten million US households. Conventional energy
sources such as fossil fuels take more time for their formation, and they are also
limited in quantity. At present, alternate energy sources need to be explored and
biofuels are considered to be an appropriate option as per the needs.
Biomass is created by green plants turning sunlight into plant material via pho-
tosynthesis, and it encompasses all land- and water-based vegetation and all organic
wastes. Photosynthesis is the process of converting chemical molecules (particu-
larly carbon dioxide and water) into food using sunlight. The majority of the chemi-
cal energy that travels through the biosphere is provided by photosynthesizing
organisms (plants, algae, and bacteria). They also produce the majority of the bio-
mass that results in the fossil fuels that influence much of today’s globe.
Photosynthesis occurs on land, at sea, and in freshwater habitats. Over 3.5 billion
years ago, the first photosynthesizing single-celled bacterium was developed
(Blankenship, 2010). Biomass is a biological substance in which the sun’s energy is
stored in chemical bonds.
Biomass residues are classified as given below. Primary residues are often formed
in the field during the cultivation of target forest products and food crops such as
maize stalks, straws, stems, and leaves. Secondary residues, on the other hand, are
formed when food crops are treated as finished goods. Agricultural and food pro-
cessing wastes include rice hulls, woodchips, palm kernel cake, coffee husk, and
sugarcane bagasse. On the other hand, tertiary residues are available after a
174 V. Ilango

Crop
residues

Forest
Sewage
residues

Biomass
sources

Industrial Animal
waste waste

Wood
waste

Fig. 3 Various sources of biomass

biomass-derived product has been expended by animals and humans and these resi-
dues may manifest as municipal solid waste (Boro et al., 2022). Various sources of
biomass are given in Fig. 3.
All organic materials derived from plants, such as crops, algae, and trees, are
referred to as biomass. The term “biomass” refers to all land- and water-based flora
as well as any organic wastes that have undergone the photosynthesis process,
which involves green plants turning sunlight into plant material. The biomass
resource may be thought of as organic waste that has chemical bonds that store solar
energy. Chemical energy is released from the substances when the bonds between
neighboring hydrogen, oxygen, and carbon molecules are broken through decom-
position, digestion, or combustion. By consuming these biological species that con-
tribute to the biomass chain, biological species will grow (Basu, 2018). In the
presence of visible light, particularly in the red region (600–700 nm) and blue
region (425–450 nm) through a process called photosynthesis, green plants release
water to get protons and electrons, which they subsequently utilize to convert car-
bon dioxide into glucose and oxygen is liberated as a waste product (Krewald et al.,
2015). Humanity has historically relied heavily on biomass as a basis of energy, and
it is currently believed that biomass accounts for between 10% and 14% of the
global energy supply (McKendry, 2002).
Carbon dioxide (CO2) is ingested by biomass during the development phase and
released after kiln ignition. In this way, biomass helps the process of recycling car-
bon dioxide (CO2) from the atmosphere and does not produce any additional green-
house gases (GHG) like CO2, methane (CH4), or nitrogen. Biomass releases the
Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 175

same amount of carbon dioxide (CO2) into the atmosphere during the development
phase as it does upon incineration, making them both “carbon-neutral” in scientific
terms (Timmons et al., 2015). Biofuels are a probable alternative transportation
option since they can be generated wherever in the globe where biomass is acces-
sible, offering an ecologically benign energy source that helps the growth of forest
agriculture, its sustainability, and linked sectors (Pishvaee et al., 2021).
By selecting certain properties, biomass production yield may be increased and
product quality can be enhanced. The following characteristics must be determined:
solid carbon content [%], particle size [nm], basic density [kg/m3], moisture absorp-
tion capacity [mg/g], moisture content [%], shape, ash, and volatile matter content
[%], calorific value [megajoules per kg fuel], heat of combustion [megajoules per kg
fuel], ash and inflammable sulfur content [%], and processing resistance and oxy-
gen, chlorine, nitrogen, and hydrogen content [%] (Rusch et al., 2021). From the
relationship between the material’s basic density and gross calorific value, energy
density is calculated according to the equation:

BED   GCV  Bd 1000

where GCV = gross calorific value [megajoules per kg fuel], Bd = basic density [kg/
m3], and BED = basic energy density [Gigajoules/m3].

3 Bioenergy Conversion Techniques

The drying process is required for the plant biomass since it is present in various
areas and variable climatic conditions and is also characterized by increasing
humidity. Before storing biomass, it is dried to eliminate moisture content and pre-
pare it for subsequent conversion operations. Shredding biomass before shipment
increases bulk density and lowers transport costs. In turn, solid fuel densification
methods (e.g., sawdust, hay, straw, and husks) are employed to recuperate energy
and physical qualities. Finally, the biomass is converted to the required standard
shapes and sizes that allow for the energy industry distribution (Kalak, 2023).
Bioenergy is the energy produced by burning biomass fuel coupled with metabolic
waste products. Bioenergy has enormous potential to satisfy future demands for
renewable energy (Ward et al., 2014).
A wide range of technologies were used and developed to generate bioenergy.
The manufacture of biofuel from waste was accomplished by the conversion of
waste energy, which may be used to supply energy. In addition to transesterification,
biochemical and thermochemical conversion processes are utilized in the produc-
tion of energy from organic waste. Biomass conversion into liquid, gaseous, and
solid forms is estimated to be resourceful and green energy providers to various
sectors.
176 V. Ilango

3.1 Thermochemical Conversion

The basic classifications of the process include hydrothermal, gasification, pyroly-


sis, combustion, and torrefaction (Nazari et al., 2021) as indicated in Fig. 4. During
this conversion, catalysts and heat are applied to the biomass so that the biopoly-
mers of the biomass are transformed into biofuels and other chemical components.
The products of this process are mainly non-condensable gases, carbon-rich solid
residue – biochar, bio-oil, and condensable vapors of bio-crude oil and tar (Daful &
Chandraratne, 2020).
Gasification, pyrolysis, and liquefaction are the three primary process options
accessible within thermochemical conversion. Research studies show that the fol-
lowing factors have led to the rising interest in thermal conversion technology: The
additional advantage of producing energy from plastic wastes cannot be digested by
microbial activity, quick processing times, minimal water use, and the ability of
industrial infrastructure to supply highly developed thermochemical transformation
equipment (Uzoejinwa et al., 2018).

3.2 Pyrolysis

Pyrolysis is the thermal decomposition of biomass reaching up to 700 °C in the


absence of oxygen, at operating temperatures ranging from 350 to 550 °C. Organic
materials are decomposed into solid, liquid, and gas mixtures during the pyrolysis
process (Lee et al., 2019). One of the most prevalent thermochemical processes
used to break down carbonaceous biomass, including cellulose, hemicellulose, and
lignin, is pyrolysis (Aravind et al., 2020). The thermal degradation of biomass
involves a number of complex reactions, including charring, dehydration, decarbox-
ylation, hydrogenation, depolymerization isomerization, and aromatization.
Typically, pyrolysis of biomass undergoes the following stages (Dabros et al., 2018):
• Heat transfer from the source to biomass to start the reaction
• Volatiles and formation of char due to the high-temperature pyrolysis of vapors
primarily

Gasification Torrefaction Combustion Hydrothermal


Pyrrolysis
Liquefaction

Fig. 4 Thermochemical conversion methods


Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 177

• Heat migration continues between fuel that is unhydrolyzed and hot volatiles,
due to the movement of hot vapors to the biomass
• Tar formation due to the volatile’s condensation related to secondary reactions
• In conjugation with primary pyrolytic reaction autocatalytic secondary pyrolysis
reactions take place

3.3 Gasification

The thermochemical oxidation process is the conversion of biomass into gas in the
presence of gasifying agents, such as air, carbon dioxide, steam, oxygen, or a mix-
ture of all these (Ruiz et al., 2013). When the generation of biogas fuel is critical,
the gasification process is performed in a low-oxygen environment to convert solid
biomass to a gaseous fuel known as synthesis gas (Cerinski et al., 2020). The final
syngas results in a mixture of methane, carbon dioxide, hydrogen, and carbon mon-
oxide and ethane, propane, and other light hydrocarbons as well as heavier hydro-
carbons such as tars. The feedstock material has an impact on the quality of the gas
generated, reactor design, presence of catalyst, reactor operational circumstances,
and gasifying agent (Parthasarathy & Narayanan, 2014). Several possible advan-
tages of biomass gasification and later conversions include provincial financial
expansion, sustainability, social and agricultural growth, and decrease in green-
house gas emissions (Demirbas & Demirbas, 2007). By overcoming the major
obstacles such as tar generation and moisture content of the biomass, the gasifica-
tion process still needs to be optimized to increase energy efficiency. New methods
have been developed to enable the efficient utilization of even hazardous and wet
biomass for energy generation.

3.4 Torrefaction

Torrefaction is a thermal pretreatment method used to change the chemical charac-


teristics of biomass in an inert atmosphere, carried out at temperatures between 200
and 300 °C (Lange, 2007). Moisture is lost during the torrefaction process, and part
of the organic components of organic compounds include hydrogen and oxygen
thermally breakdown, producing volatile organic chemicals (Beckman et al., 2011).
Depending on the degree of torrefaction, the residual solid biomass (solid) con-
tains roughly 30% more energy per unit of the mass. The properties of biomass are
equivalent to coal at a temperature between 275 and 300 °C, an inert atmosphere, a
residence time of 20 to 40 min, and a heating rate of 10 °C/min to maintain the reac-
tor temperature (Mamvura et al., 2018). To determine the kinetic and drying models
of the torrefaction process, substantial wet wood particle was used (Basu et al.,
178 V. Ilango

2014). They obtained residence time, optimum temperature, and particle size as
10 h, 290 °C, and 24 mm, respectively (Kamila et al., 2017). The major fiber com-
positions in common plants and biomass are lignin, cellulose, and hemicellulose.
They are quite different in structural, chemical, and physical properties. Different
reactions occur, and all these three components are decomposed as the biomass is
torrefied. At a temperature less than 250 °C, hemicellulose was dried and decom-
posed while cellulose and lignin were not. The volatiles and organic polymers in
cellulose and lignin were removed or decomposed as the temperature increased
above 250 °C (Chen & Anderson, 1980).

3.5 Combustion

Power generation of biomass by direct combustion has continued since the 1990s
(Yin et al., 2008). Condensing steam turbine, electric generator, and vibrating grate
boiler are all components of biomass-fired combined heat and power systems (Chen
et al., 2021). There are four different stages in combustion: drying, pyrolysis (de-­
volatilization), volatile combustion, and char combustion. The process results
mainly depend on particle size, combustion environment, the characteristics of the
feedstock, and temperature, which can contain char and ash as the solid by-products
of combustion (usually comprised of inorganic oxides and carbonates) (Adams
et al., 2018). On complete combustion of biomass, water vapor, and carbon dioxide
are produced. The major methods for reducing the problems include combustion
process modifications, such as boiler modifications, flue gas recirculation, and re-­
burning technologies, which frequently reduce such emissions at a reasonable cost
(Oluwoye et al., 2020).

3.6 Hydrothermal Liquefaction

When biomass is subjected to temperatures between 250 and 400 °C, extremely
high pressure (4 to 25 MPa) (Huber et al., 2006) with the application of catalysts
(mostly alkali) for an extended period of time, a thermochemical process known as
hydrous pyrolysis or hydrothermal liquefaction (HTL) occurs. It uses subcritical
water (SCW) and breaks down the biomass biopolymeric structure to primarily liq-
uid components, producing a liquid fuel known as “biocrude” or bio-oil from pyrol-
ysis, and biocrude is physically much less dense and much more viscous (Dimitriadis
& Bezergianni, 2017). Several intricate mechanisms, such as hydrolysis, are acti-
vated during the HTL process, degrading biomass macromolecules before breaking
them down into reactive fragments through bond cleavage and a number of reac-
tions, including deoxygenation, decarboxylation, dehydration, and dehydrogena-
tion. Meanwhile, some intricate chemicals, such as biocrude, are produced through
Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 179

depolymerization (Devi & Parthiban, 2020). HTL produces biocrude from organic
matter under certain states, such as the existence of water in hydrothermal condi-
tions. The water either remains highly dense supercritical or liquid under these cir-
cumstances (Seehar et al., 2020).

3.7 Biochemical Conversion

Biochemical conversion is using specific bacteria yeast or yeast to transform waste


or biomass into usable energy. The traditional process choices for producing various
biofuels include anaerobic digestion, alcoholic fermentation, and photobiological
methods (Lee et al., 2019). The method offers a platform for the production of fuels
and chemicals such as ethanol, biogas, hydrogen, acetone, butanol, and a variety of
organic acids (Garba, 2021).

3.8 Anaerobic Digestion

Management issues are being brought on by the ongoing garbage output. Traditional
waste management practices such as landfilling and incineration generate gases that
could contribute to global warming. Due to rapid population growth and industrial-
ization, there is likewise an exponential rise in energy demands. In order to encoun-
ter this continuously rising energy demand, access to green and clean energy is
crucial for the sustainable development of human society. These two problems can
complement one another’s remedies if they are handled properly using biowaste-to-­
bioenergy (BtB) technology (Bhatia et al., 2018). Along with thermochemical oper-
ation, biochemical methods such as anaerobic digestion (AD) and fermentation
offer hope for the future of renewable energy sources (Batista et al., 2018). Anaerobic
digestion of the biomass waste left over after the biodiesel synthesis process is one
method for maximizing nutritional extraction.
Microorganisms convert waste microalgae biomass into biogas during anaero-
bic digestion. The biogas produced is mostly composed of CO2 and CH4, with
trace amounts of hydrogen sulfide. Biogas produced has a lower heating value and
energy content of 20–40% of the biomass. A moisture content of up to 90% wet
biomass can be utilized in anaerobic digestion (Brennan & Owende, 2010).
Important factors that affect the production of biogas include pH, the absence of
oxygen in the interaction environment, the density, type, and raw material content,
and the amount of time needed to halt the fluid inside the reservoir (Bijarchiyan
et al., 2020).
180 V. Ilango

3.9 Alcoholic Fermentation

The fermentation process for wines and ciders is usually carried out using strains of
Saccharomyces cerevisiae, the most prevalent and readily accessible yeast
(Puligundla et al., 2011). The objective of biomass fermentation is to biologically
convert biomass into chemicals, materials, combustible gases, or fuels such as bio-
ethanol and biobutanol. The chemical equation for the production of ethanol from
glucose (Maicas, 2020) is as follows:

Glucose C6 H12 O6  Ethanol 2C2 H 5OH  Carbon dioxide CO2

When starch or molasses are used as feedstocks, the simple sugars (such as glu-
cose, sucrose, and carbohydrate-rich biowaste) are fermented into bioethanol or
biobutanol. The primary operational factors affecting ethanol output and fermenta-
tion effectiveness are the presence and removal of many by-products, osmotic pres-
sure (toxic to yeasts), and yeast metabolism, including hydroxymethyl furfural
(Manikandan et al., 2023). A computerized forecast of biomethane manufacture to
examine a specific application was developed by a combination of specified organic
waste using STATISTICA 10 software and a two-factor central composite design
technique. A concurrent laboratory-scale fermentation experiment was conducted to
check the predicted results, and the results demonstrated the use of computer tech-
nology in this particular application with a greater level of consistency (Mateescu
et al., 2022).

3.10 Photobiological Methods

In the presence of light, specific biomass, such as microalgae, naturally generate


hydrogen gas. Enzyme hydrogenase in an anaerobic environment subsequently con-
verts the hydrogen ions to hydrogen gas. The hydrogenase enzymes are rapidly
repressed by the oxygen produced during photosynthesis, resulting in the termina-
tion of hydrogen gas production. This shows that cultivating microalgae to produce
H2 gas requires anaerobic conditions (Cantrell et al., 2008). Another method
involves using a two-stage system, where the first stage involves the growth of
microalgae under normal conditions and the second stage involves encouraging
constant H2 production under sulfur-deficient and anaerobic environments (Ghirardi,
2000). For the purpose of producing ethanol on an industrial scale, some plants with
a high content of saccharides and starch are purposefully cultivated. These include
crops such as sugarcane, potatoes, or grains. This waste biomass-derived bioethanol
can be added to gasoline or diesel as a fuel additive (Chavan et al., 2022).
Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 181

3.11 Advantages and Limitations of Bioenergy Production


Methods (Table 1)

Table 1 Advantages and limitations of bioenergy production methods


Methods Advantages Limitations References
Thermal conversion methods
1. Pyrolysis Energy and time savings, Chemical instability of the Motasemi and
less greenhouse gas and oil, biochar, one of the key Afzal (2013), and
carbon emissions products is of less industrial Kundu et al.
value (2017)
2. Gasification Ideal route for the The melting of mineral Santos et al.
conversion of diverse species caused by the usage (2022)
biomass feedstocks of high temperatures
resulted in defluidization
and bed agglomeration
3. Torrefication Loss in weight of the Despite its better potential, Shankar
biomass as volatiles, torrefaction technology has Tumuluru et al.
shrinks, becomes fragile, been slow to commercialize (2011), and
makes it easier to grind and due to the difficulties with Tumuluru et al.
pulverize reactor design and end (2021)
product quality
4. Combustion Net carbon emissions can Biomass fuel supply is Sheth and Babu
be greatly decreased, limited to particularly (2010), and Varol
resulting in a clean cost-effective biomass et al. (2018)
development mechanism sources derived from waste
and reducing greenhouse wood products or
gas (GHG) emissions agro-processing processes
5. Hydrothermal Feedstock with a high A major issue for running a Kaur et al.
liquefaction moisture content without full-scale facility is feeding (2020), and
drying can be used biomass feedstock into the Chandraratne and
reactor at high pressure Daful (2022)
6. Alcoholic Hydrolysis methods utilize Acidic environment may Lee et al. (2019)
fermentation enzymes and acid/alkali. alter the sugars into
Acid treatment is fast and undesirable forms
cheap
7. Anaerobic Reduces pollutant levels, H2S and NH3 are harmful Shrestha et al.
digestion stabilizes sludge, and and highly corrosive, (2020), and
reduces sludge tonnage causing harm to combined Uddin et al.
significantly with minimal heat and power units and (2021)
energy input metal parts
8. Photobiological The most efficient and Though hydrogen Chavan et al.
methods environmentally friendly productivity is high due to (2022), and
way of creating hydrogen, the dependence on the sun Gaweł Sołowski
with low-to-zero carbon it is unstable (2018)
emissions
182 V. Ilango

4 Challenges Faced

Global climate change and greenhouse gas emissions are contentious problems that
have a significant impact on society, the energy sector, and government policy
(Kalicki & Goldwyn, 2005). The use of a variety of sustainable and renewable fuels
is necessary for the enduring economic viability of the conveyance and energy in
the industrial sectors (Chung, 2014). Such energy sources ought to have high energy
contents and low greenhouse gas emissions. It is crucial that the techniques used to
develop the resources and produce these fuels have little effect on the environment,
water supply, food chain, and land use. Since they are almost carbon-neutral and
wind, nuclear, geothermal, solar, and biomass are all great sources of energy, high
moisture content biomass products are not appropriate as fuel for traditional ther-
mochemical conversion processes such as pyrolysis and gasification. High moisture
levels can make conversion processes less effective.
Because moisture in raw biomass sources is also unbearable, fuel made from
these materials may contain moisture. The fuels cannot burn quickly since they
contain a lot of moisture. Water that is contained in the fuel is vaporized using a
portion of the energy in the fuel (Irmak, 2019).
Bioenergy can be competitive in some instances, especially when biomass fuel,
such as garbage and leftovers, is inexpensive or even free. While a number of bio-
energy technologies are nearing maturity, more technological development is
required to enable promising technologies to reach commercial production and
attain cost competitiveness. Many phases are involved in biochemical fermentation
conversion processes, including biomass fractionation and pretreatment, microbial
fermentation, saccharification, enzymatic hydrolysis, product separation, and puri-
fication. Research is required to significantly improve the efficacy of each of these
inherent processes. The difficulties are connected to the accessibility of raw materi-
als in suitable amounts and at competitive costs (Virkajärvi et al., 2009). Anaerobic
digestion also has the potential to make a substantial influence on future biofuel and
bioenergy generation.
The exploration for cost-effective, skilled integrated systems to produce biofer-
tilizers and bioenergy (methane) from organic feedstocks, such as microalgal bio-
mass, crop residues, purpose-grown energy crops, animal and food waste, and
additional organic materials, should be continued to materialize effective technolo-
gies in the future. The energy crisis, greenhouse gas emissions, and air pollution
have accelerated the evolution to renewable and cleaner fuel sources. Biomass is a
great prospective fuel source since its resource growth and manufacturing methods
have a diminutive influence on the water supply, land usage, food chain, and envi-
ronment (Chung, 2013).
Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 183

5 Future Prospects and Alternatives

Industry will surely continue to expand, but the rate and expansion routes, as well
as the repercussions on the people and environment, are impossible to predict.
Despite all of these unknowables, it is obvious that industrial and agricultural waste
will continue to be produced, necessitating proper and efficient management.
Recent study has shown that utilizing biomass waste for energy purposes is a prom-
ising and novel development route, employing environmentally friendly methods
that limit hazardous material emissions and are carbon dioxide neutral. Sawdust,
perennial grasses, branches, landfill gases, and post-production waste are just a few
examples of biomass resources that may be exploited to generate liquid fuels, heat,
and energy (Kalak, 2023).
According to international forecasts, by 2050, the world population will need
50% additional water, 70% extra food, a 50–80% decrease in carbon dioxide (CO2)
emissions, and 50% added fuel to maintain political, climate, and social security.
Presently, fossil fuels satisfy approximately 80% of the world’s energy requirement
and are responsible for a significant portion of greenhouse gas emissions (Yang
et al., 2016; Mathimani et al., 2019). It is critical to evoke that biomass markets will
increase the value of biomass products, residues, and productive lands. This value
will assist to increase the economic sustainability of working lands and will operate
as a positive incentive to help preserve farms and forests from the looming danger
of urban and suburban sprawl – the most significant land use effect (Dorsey, 2019).
Through the execution of biorefinery platforms and the approximation of low
environmental impact technologies and processes, new sustainable production strat-
egies for biofuels and energy can be recognized, making these biobased industries
more modest alternatives and revealing the economy of existing value chains
(Clauser et al., 2021). The characteristics of the biomass feedstock have a direct
impact on the selection of conversion routes and bioenergy end products, which in
turn has an impact on the synthesis of waste biorefinery. As a result, the majority of
waste biorefineries are categorized based on the kind or source of feedstocks. As a
result, merging waste biorefineries to process different biomass wastes as feed-
stocks demands a technically and economically feasible aggregation method
(Ochieng et al., 2022).

6 Conclusions

Biomass, a renewable and carbon-neutral energy source, has a significant potential


to meet our future energy demands. Biomass resources face various hurdles before
they can be completely used for biofuel and other valuable products. The availabil-
ity, abundance, and needs for expansion rate and other criteria have a significant
184 V. Ilango

impact on feedstock choices for value-added goods. The current level of energy and
biofuel manufacture from biomass demonstrates that the use of biorefinery plat-
forms is a viable production approach. The future of biorefineries seems bright. To
fully realize their potential, governments and international organizations must direct
technical advancements and public policies toward high-value goods from biomass
and the development of fuels.
The characteristics of the biomass feedstock have a direct impact on the selection
of conversion routes and bioenergy end products, which in turn has an impact on the
synthesis of a waste biorefinery. As a result, the majority of waste biorefineries are
categorized based on the kind or source of feedstocks. As a result, merging waste
biorefineries to process different biomass wastes as feedstocks demands a techni-
cally and economically feasible aggregation method.
Their use not only alleviates waste issues and promotes alternative energy
sources, but it also draws attention to the hot issue of bioenergy with carbon capture
and storage (BECCS). Biofuels derived from biomass waste are still thought to be
more robust in terms of transportation, material handling, and conversion technol-
ogy than typical editable food crop-based biofuels. As humanity strives to reduce its
dependency on fossil fuel-based energy, biomass energy is set to become a future
alternative energy source. This compilation provides a thorough understanding of
biomass, its production techniques, challenges, and future prospects, and it can be
utilized efficiently in the future by any individual for in-depth research and simple
reference.

References

Adams, P., Bridgwater, T., Lea-Langton, A., Ross, A., & Watson, I. (2018). Chapter 8 – Biomass
conversion technologies. In [online] ScienceDirect. Available at: https://linkinghub.elsevier.
com/retrieve/pii/B9780081010365000082. Accessed July 14, 2023.
Adekoya, O. B., Akinbayo, S. B., Ishola, O. A., & Abdulaziz, M. (2023). Are all the U.S. biomass
energy sources green? Energy Policy, 179, 113614. https://doi.org/10.1016/j.enpol.2023.113614
Aravind, S., Kumar, P. S., Kumar, N. S., & Siddarth, N. (2020). Conversion of green algal bio-
mass into bioenergy by pyrolysis. A review. Environmental Chemistry Letters, 18(3), 829–849.
https://doi.org/10.1007/s10311-­020-­00990-­2
Basu, P. (2018). Biomass gasification, pyrolysis and torrefaction: Practical design and theory. In
[online] Google Books. Academic Press. Available at: https://books.google.co.in/books?hl=en
&lr=&id=BYM2DwAAQBAJ&oi=fnd&pg=PP1&ots=nItDd9vCkJ&sig=TbengDsoitQVcBX
VWALukbohAPA&redir_esc=y#v=onepage&q&f=false. Accessed July 17, 2023.
Basu, P., Sadhukhan, A. K., Gupta, P., Rao, S., Dhungana, A., & Acharya, B. (2014). An experi-
mental and theoretical investigation on torrefaction of a large wet wood particle. Bioresource
Technology, 159, 215–222. https://doi.org/10.1016/j.biortech.2014.02.105
Batista, A. P., Gouveia, L., & Marques, P. A. S. S. (2018). Fermentative hydrogen production
from microalgal biomass by a single strain of bacterium Enterobacter aerogenes – Effect of
operational conditions and fermentation kinetics. Renewable Energy, 119, 203–209. https://
doi.org/10.1016/j.renene.2017.12.017
Beckman, J., Hertel, T., Taheripour, F., & Tyner, W. (2011). Structural change in the biofuels
era. European Review of Agricultural Economics, 39(1), 137–156. https://doi.org/10.1093/
erae/jbr041
Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 185

Bhatia, S. K., Joo, H.-S., & Yang, Y.-H. (2018). Biowaste-to-bioenergy using biological methods –
A mini-review. Energy Conversion and Management, 177, 640–660. https://doi.org/10.1016/j.
enconman.2018.09.090
Bijarchiyan, M., Sahebi, H., & Mirzamohammadi, S. (2020). A sustainable biomass network
design model for bioenergy production by anaerobic digestion technology: Using agricul-
tural residues and livestock manure. Energy, Sustainability and Society, 10(1). https://doi.
org/10.1186/s13705-­020-­00252-­7
Blankenship, R. E. (2010). Early evolution of photosynthesis. Plant Physiology, [online], 154(2),
434–438. https://doi.org/10.1104/pp.110.161687
Boro, M., Verma, A. K., Chettri, D., Yata, V. K., & Verma, A. K. (2022). Strategies involved in
biofuel production from agro-based lignocellulose biomass. Environmental Technology &
Innovation, [online], 102679. https://doi.org/10.1016/j.eti.2022.102679
Brennan, L., & Owende, P. (2010). Biofuels from microalgae—A review of technologies for pro-
duction, processing, and extractions of biofuels and co-products. Renewable and Sustainable
Energy Reviews, 14(2), 557–577. https://doi.org/10.1016/j.rser.2009.10.009
Cantrell, K. B., Ducey, T., Ro, K. S., & Hunt, P. G. (2008). Livestock waste-to-bioenergy gen-
eration opportunities. Bioresource Technology, [online], 99(17), 7941–7953. https://doi.
org/10.1016/j.biortech.2008.02.061
Center for Sustainable Systems. (2021). U.S. Renewable Energy Factsheet. [online]. Available at:
https://css.umich.edu/publications/factsheets/energy/us-­renewable-­energy-­factsheet
Cerinski, D., Baleta, J., Mikulčić, H., Mikulandrić, R., & Wang, J. (2020). Dynamic modelling of
the biomass gasification process in a fixed bed reactor by using the artificial neural network.
Cleaner Engineering and Technology, 1, 100029. https://doi.org/10.1016/j.clet.2020.100029
Chandraratne, M. R., & Daful, G. A. (2022). Recent advances in thermochemical conversion of bio-
mass. In Recent perspectives in pyrolysis research. https://doi.org/10.5772/intechopen.100060
Chaudry, S., Bahri, P. A., & Moheimani, N. R. (2015). Pathways of processing of wet microalgae
for liquid fuel production: A critical review. Renewable and Sustainable Energy Reviews, 52,
1240–1250. https://doi.org/10.1016/j.rser.2015.08.005
Chavan, N., Bhor, D., & Pote, R. (2022). A short review on biomass production from agricultural
wastes. International Journal of Engineering Research & Technology, [online], 11(6). https://
doi.org/10.17577/IJERTV11IS060041
Chen, W., & Anderson, A. S. (1980). Extraction of hemicellulose from ryegrass straw for the
production of glucose isomerase and use of the resulting straw residue for animal feed, 22(3),
519–531. https://doi.org/10.1002/bit.260220305
Chen, H., Xue, K., Wu, Y., Xu, G., Jin, X., & Liu, W. (2021). Thermodynamic and economic
analyses of a solar-aided biomass-fired combined heat and power system. Energy, 214, 119023.
https://doi.org/10.1016/j.energy.2020.119023
Chung, J. N. (2013). Grand challenges in bioenergy and biofuel research: Engineering and technol-
ogy development, environmental impact, and sustainability. Frontiers in Energy Research, 1.
https://doi.org/10.3389/fenrg.2013.00004
Chung, J. N. (2014). A theoretical study of two novel concept systems for maximum thermal-­
chemical conversion of biomass to hydrogen. Frontiers in Energy Research, 1. https://doi.
org/10.3389/fenrg.2013.00012
Clauser, N. M., González, G., Mendieta, C. M., Kruyeniski, J., Area, M. C., & Vallejos,
M. E. (2021). Biomass waste as sustainable raw material for energy and fuels. Sustainability,
13(2), 794. https://doi.org/10.3390/su13020794
Dabros, T. M. H., Stummann, M. Z., Høj, M., Jensen, P. A., Grunwaldt, J.-D., Gabrielsen, J.,
Mortensen, P. M., & Jensen, A. D. (2018). Transportation fuels from biomass fast pyroly-
sis, catalytic hydrodeoxygenation, and catalytic fast hydropyrolysis. Progress in Energy and
Combustion Science, [online], 68, 268–309. https://doi.org/10.1016/j.pecs.2018.05.002
Daful, A. G., & Chandraratne, M. R. (2020). Biochar production from biomass waste-derived
material. In [online] ScienceDirect. Available at: https://linkinghub.elsevier.com/retrieve/pii/
B9780128035818112494. Accessed July 10, 2023.
186 V. Ilango

de Carraro, C. F. F., Martins, A. C., da Faria, A. C. S., & Loures, C. C. A. (2020). Agroenergy from
residual biomass: Energy perspective. [online] www.intechopen.com. Available at: https://
www.intechopen.com/chapters/74065. Accessed August 13, 2023.
Demirbas, A. H., & Demirbas, I. (2007). Importance of rural bioenergy for developing coun-
tries. Energy Conversion and Management, 48(8), 2386–2398. https://doi.org/10.1016/j.
enconman.2007.03.005
Devi, T. E., & Parthiban, R. (2020). Hydrothermal liquefaction of Nostoc ellipsosporum biomass
grown in municipal wastewater under optimized conditions for bio-oil production. Bioresource
Technology, 316, 123943. https://doi.org/10.1016/j.biortech.2020.123943
Dimitriadis, A., & Bezergianni, S. (2017). Hydrothermal liquefaction of various biomass and
waste feedstocks for biocrude production: A state of the art review. Renewable and Sustainable
Energy Reviews, 68, 113–125. https://doi.org/10.1016/j.rser.2016.09.120
Dorsey, P. (2019). Bioenergy (Biofuels and Biomass) | EESI. [online] Eesi.org. Available at:
https://www.eesi.org/topics/bioenergy-­biofuels-­biomass/description
Garba, A. (2021). Biomass conversion technologies for bioenergy generation: An introduction. In
Biotechnological applications of biomass. https://doi.org/10.5772/intechopen.93669
Ghirardi, M. (2000). Microalgae: A green source of renewable H2. Trends in Biotechnology,
18(12), 506–511. https://doi.org/10.1016/s0167-­7799(00)01511-­0
Huber, G. W., Iborra, S., & Corma, A. (2006). Synthesis of transportation fuels from biomass:
Chemistry, catalysts, and engineering. Chemical Reviews, 106(9), 4044–4098. https://doi.
org/10.1021/cr068360d
Irmak, S. (2019). Challenges of biomass utilization for biofuels. In Biomass for bioenergy – Recent
trends and future challenges. [online]. https://doi.org/10.5772/intechopen.83752
Kalak, T. (2023). Potential use of industrial biomass waste as a sustainable energy source in the
future. Energies, 16(4), 1783. https://doi.org/10.3390/en16041783
Kalicki, J. H., & Goldwyn, D. L. (2005). Energy and security: Toward a new foreign policy strat-
egy. In [online] Google Books. Woodrow Wilson Center Press. Available at: https://books.
google.co.in/books/about/Energy_and_Security.html?id=IWO0AAAAIAAJ&redir_esc=y.
Accessed July 26, 2023.
Kamila, B., Sadhukhan, A. K., Gupta, P., Basu, P., Regmi, B., Dutta, A., & Acharya, B. (2017).
Two-dimensional modeling of torrefaction of a large biomass particle. International Journal of
Green Energy, 14(13), 1119–1129. https://doi.org/10.1080/15435075.2017.1359785
Kaur, R., Biswas, B., Kumar, J., Jha, M. K., & Bhaskar, T. (2020). Catalytic hydrothermal lique-
faction of castor residue to bio-oil: Effect of alkali catalysts and optimization study. Industrial
Crops and Products, 149, 112359–112359. https://doi.org/10.1016/j.indcrop.2020.112359
Krewald, V., Retegan, M., & Pantazis, D. A. (2015). Principles of natural photosynthesis. Topics in
Current Chemistry, 23–48. https://doi.org/10.1007/128_2015_645
Kumar, M. (2020). Social, economic, and environmental impacts of renewable energy resources.
[online] www.intechopen.com. IntechOpen. Available at: https://www.intechopen.com/
chapters/70874
Kundu, K., Chatterjee, A., Bhattacharyya, T., Roy, M., & Kaur, A. (2017). Thermochemical con-
version of biomass to bioenergy: A review. In Prospects of alternative transportation fuels
(pp. 235–268). https://doi.org/10.1007/978-­981-­10-­7518-­6_11
Lange, J.-P. (2007). Lignocellulose conversion: An introduction to chemistry, process and econom-
ics. Biofuels, Bioproducts and Biorefining, 1(1), 39–48. https://doi.org/10.1002/bbb.7
Lee, S. Y., Sankaran, R., Chew, K. W., Tan, C. H., Krishnamoorthy, R., Chu, D.-T., & Show,
P.-L. (2019). Waste to bioenergy: A review on the recent conversion technologies. BMC Energy,
1(1). https://doi.org/10.1186/s42500-­019-­0004-­7
Maicas, S. (2020). The role of yeasts in fermentation processes. Microorganisms, [online], 8(8),
1142. https://doi.org/10.3390/microorganisms8081142
Mamvura, T. A., Pahla, G., & Muzenda, E. (2018). Torrefaction of waste biomass for application in
energy production in South Africa. South African Journal of Chemical Engineering, 25, 1–12.
https://doi.org/10.1016/j.sajce.2017.11.003
Bioenergy Production Using Biomass Wastes: Challenges of Circular Economy 187

Manikandan, S., Vickram, S., Sirohi, R., Subbaiya, R., Krishnan, R. Y., Karmegam, N.,
Sumathijones, C., Rajagopal, R., Chang, S. W., Ravindran, B., & Awasthi, M. K. (2023).
Critical review of biochemical pathways to transformation of waste and biomass into bio-
energy. Bioresource Technology, 372, 128679. https://doi.org/10.1016/j.biortech.2023.128679
Mateescu, C., Tudor, E., Dima, A.-D., Chirita, I. C., Tanasiev, V., & Prisecaru, T. (2022). Artificial
intelligence approach in predicting biomass-to-biofuels conversion performances. https://doi.
org/10.1109/iccc54292.2022.9805871
Mathimani, T., & Mallick, N. (2019). A review on the hydrothermal processing of microalgal bio-
mass to bio-oil – Knowledge gaps and recent advances. Journal of Cleaner Production, 217,
69–84. https://doi.org/10.1016/j.jclepro.2019.01.129
Mathimani, T., Baldinelli, A., Rajendran, K., Prabakar, D., Matheswaran, M., Pieter van Leeuwen,
R., & Pugazhendhi, A. (2019). Review on cultivation and thermochemical conversion of
microalgae to fuels and chemicals: Process evaluation and knowledge gaps. Journal of Cleaner
Production, [online], 208, 1053–1064. https://doi.org/10.1016/j.jclepro.2018.10.096
McKendry, P. (2002). Energy production from biomass (part 1): Overview of biomass. Bioresource
Technology, 83(1), 37–46. https://doi.org/10.1016/s0960-­8524(01)00118-­3
Motasemi, F., & Afzal, M. T. (2013). A review on the microwave-assisted pyrolysis tech-
nique. Renewable and Sustainable Energy Reviews, 28, 317–330. https://doi.org/10.1016/j.
rser.2013.08.008
Nair, L. G., Agrawal, K., & Verma, P. (2022). An overview of sustainable approaches for bioenergy
production from agro-industrial wastes. Energy Nexus, 6, 100086. https://doi.org/10.1016/j.
nexus.2022.100086
Nazari, L., Xu, C. (C.)., & Ray, M. B. (2021). Advanced technologies (biological and thermochemi-
cal) for waste-to-energy conversion. In Advanced and emerging technologies for resource recov-
ery from wastes (pp. 55–95). Springer Singapore. https://doi.org/10.1007/978-­981-­15-­9267-­6_3
Ochieng, R., Gebremedhin, A., & Sarker, S. (2022). Integration of waste to bioenergy conversion
systems: A critical review. Energies, 15(7), 2697. https://doi.org/10.3390/en15072697
Oluwoye, I., Altarawneh, M., Gore, J., & Dlugogorski, B. Z. (2020). Products of incomplete com-
bustion from biomass reburning. Fuel, 274, 117805. https://doi.org/10.1016/j.fuel.2020.117805
Parthasarathy, P., & Narayanan, K. S. (2014). Hydrogen production from steam gasification of
biomass: Influence of process parameters on hydrogen yield – A review. Renewable Energy,
[online], 66, 570–579. https://doi.org/10.1016/j.renene.2013.12.025
Pishvaee, M. S., Mohseni, S., & Bairamzadeh, S. (2021). An overview of biomass feedstocks for
biofuel production. In Biomass to biofuel supply chain design and planning under uncertainty.
https://doi.org/10.1016/b978-­0-­12-­820640-­9.00001-­5
Puligundla, P., Smogrovicova, D., Obulam, V. S. R., & Ko, S. (2011). Very high gravity (VHG)
ethanolic brewing and fermentation: A research update. Journal of Industrial Microbiology &
Biotechnology, 38(9), 1133–1144. https://doi.org/10.1007/s10295-­011-­0999-­3
Ruiz, J. A., Juárez, M. C., Morales, M. P., Muñoz, P., & Mendívil, M. A. (2013). Biomass gas-
ification for electricity generation: Review of current technology barriers. Renewable and
Sustainable Energy Reviews, 18, 174–183. https://doi.org/10.1016/j.rser.2012.10.021
Rusch, F., Raul, D., & Hillig, É. (2021). Energy properties of bamboo biomass and mate co-­
products. SN Applied Sciences, 3(6). https://doi.org/10.1007/s42452-­021-­04584-­7
Santos, S. M., Assis, A. C., Gomes, L., Nobre, C., & Brito, P. (2022). Waste gasification technolo-
gies: A brief overview. Waste, 1(1), 140–165. https://doi.org/10.3390/waste1010011
Seehar, T. H., Toor, S. S., Shah, A. A., Pedersen, T. H., & Rosendahl, L. A. (2020). Biocrude pro-
duction from wheat straw at sub and supercritical hydrothermal liquefaction. Energies, 13(12),
3114. https://doi.org/10.3390/en13123114
Shankar Tumuluru, J., Sokhansanj, S., Hess, J. R., Wright, C. T., & Boardman, R. D. (2011).
REVIEW: A review on biomass torrefaction process and product properties for energy applica-
tions. Industrial Biotechnology, 7(5), 384–401. https://doi.org/10.1089/ind.2011.7.384
Sheth, P. N., & Babu, B. V. (2010). Production of hydrogen energy through biomass (waste wood)
gasification. International Journal of Hydrogen Energy, 35(19), 10803–10810. https://doi.
org/10.1016/j.ijhydene.2010.03.009
188 V. Ilango

Shrestha, B., Hernandez, R., Fortela, D. L. B., Sharp, W., Chistoserdov, A., Gang, D., Revellame,
E., Holmes, W., & Zappi, M. E. (2020). A review of pretreatment methods to enhance sol-
ids reduction during anaerobic digestion of municipal wastewater sludges and the resulting
digester performance: Implications to future urban biorefineries. Applied Sciences, [online],
10(24), 9141. https://doi.org/10.3390/app10249141
Sołowski, G. (2018). Biohydrogen production – Sources and methods: A review. International
Journal of Bioprocessing and Biotechniques. [online]. Available at: https://www.gavinpublish-
ers.com/article/view/biohydrogen-­production-­sources-­and-­methods-­a-­review
Speight, J. G., & Singh, K. (2014). Environmental management of energy from biofuels and bio-
feedstocks. John Wiley & Sons, Inc.. https://doi.org/10.1002/9781118915141
Timmons, D. S., Buchholz, T., & Veeneman, C. H. (2015). Forest biomass energy: Assessing atmo-
spheric carbon impacts by discounting future carbon flows. GCB Bioenergy, 8(3), 631–643.
https://doi.org/10.1111/gcbb.12276
Tumuluru, J. S., Ghiasi, B., Soelberg, N. R., & Sokhansanj, S. (2021). Biomass torrefaction pro-
cess, product properties, reactor types, and moving bed reactor design concepts. Frontiers in
Energy Research, 9. https://doi.org/10.3389/fenrg.2021.728140
Uddin, M. N., Siddiki, S. Y. A., Mofijur, M., Djavanroodi, F., Hazrat, M. A., Show, P. L., Ahmed,
S. F., & Chu, Y.-M. (2021). Prospects of bioenergy production from organic waste using
anaerobic digestion technology: A mini review. Frontiers in Energy Research, 9. https://doi.
org/10.3389/fenrg.2021.627093
Uzoejinwa, B. B., He, X., Wang, S., El-Fatah Abomohra, A., Hu, Y., & Wang, Q. (2018).
Co-pyrolysis of biomass and waste plastics as a thermochemical conversion technology for high-­
grade biofuel production: Recent progress and future directions elsewhere worldwide. Energy
Conversion and Management, 163, 468–492. https://doi.org/10.1016/j.enconman.2018.02.004
Varol, M., Symonds, R., Anthony, E. J., Lu, D., Jia, L., & Tan, Y. (2018). Emissions from co-firing
lignite and biomass in an oxy-fired CFBC. Fuel Processing Technology, 173, 126–133. https://
doi.org/10.1016/j.fuproc.2018.01.002
Virkajärvi, I., Niemelä, M. V., Hasanen, A., & Teir, A. (2009). Cellulosic ethanol via biochemi-
cal processing poses a challenge for developers and implementors. BioResources, [online],
4(4), 1718. Available at: https://www.academia.edu/82842054/Cellulosic_Ethanol_via_
Biochemical_Processing_Poses_a_Challenge_for_Developers_and_Implementors. Accessed
August 10, 2023
Ward, J., Rasul, M. G., & Bhuiya, M. M. K. (2014). Energy recovery from biomass by fast pyroly-
sis. Procedia Engineering, 90, 669–674. https://doi.org/10.1016/j.proeng.2014.11.791
Yang, I.-S., Salama, E.-S., Kim, J.-O., Govindwar, S. P., Kurade, M. B., Lee, M., Roh, H.-S., &
Jeon, B.-H. (2016). Cultivation and harvesting of microalgae in photobioreactor for biodiesel
production and simultaneous nutrient removal. Energy Conversion and Management, 117,
54–62. https://doi.org/10.1016/j.enconman.2016.03.017
Yin, C., Rosendahl, L. A., & Kær, S. K. (2008). Grate-firing of biomass for heat and power produc-
tion. Progress in Energy and Combustion Science, 34(6), 725–754. https://doi.org/10.1016/j.
pecs.2008.05.002
Application of Enzymes in Biomass Waste
Management

Preeti Ranjan, Maneesh Kumar, Himanshu Bhardwaj, Priyanka Kumari,


and Arti Kumari

1 Introduction

The intensification of agricultural and processing operations is a consequence of the


increasing demand for food and other vital resources due to population growth. It
leads to the production of biomass waste. Intensive agricultural practices often
involve high livestock densities or large-scale crop production. These practices aim
to maximize production and efficiency, but they can also generate significant
amounts of waste (Rasul & Sharma, 2016). Agricultural waste disposal, recycling,
and management are not practiced everywhere, leading to a growing problem. In
many developing countries, a significant portion of biomass residues are either left
to rot in the field or burned in the open, causing significant environmental damage.
Rapid global urbanization and increasing demand for building products also require
alternative, sustainable energy sources and raw material supplies (Karlsson et al.,
2020). The potential of biomass waste as a valuable resource for energy and mate-
rial production has not yet been fully realized, and too little attention is paid to
developing a low-carbon approach to its use. In order to get the most out of the huge

P. Ranjan
Department of Biotechnology, Patliputra University, Patna, India
M. Kumar (*)
Department of Biotechnology, Magadh University, Bodh Gaya, Bihar, India
H. Bhardwaj
Department of Biochemistry, Patna University, Patna, Bihar, India
P. Kumari
Department of Biotechnology, Gautam Buddha University,
Greater Noida, Uttar Pradesh, India
A. Kumari
Department of Biotechnology, Patna Women’s College, Patna, Bihar, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 189
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_10
190 P. Ranjan et al.

biomass waste, many biomass waste management methods and capabilities have
been developed. Biomass waste management is about environmentally sound dis-
posal and treatment. Biomass waste is an increasing source of carbon-based materi-
als due to its high carbon content, low cost, accessibility, ubiquity, renewability, and
environmental friendliness (Soffian et al., 2022). Organic materials include things
such as food waste, agricultural residues, and forestry residues, as well as energy-­
specific crops and aquatic biomass. These include the stalks, husks, leaves, and
stems that are left over after harvesting maize, wheat, rice, or sugarcane. Straw and
bagasse are examples of waste products derived from biomass (Antero et al., 2020).
Any edible substance left behind during food production, processing, distribution,
and consumption is food waste. This includes spoiled or expired food, leftovers,
peels, and leftovers from households, restaurants, and the food industry (Secondi
et al., 2019). The wood waste consists of twigs, tops, bark, and sawdust. These com-
ponents are generated during timber harvesting, wood processing, and forestry
operations (Zyryanov et al., 2021). Special energy crops are being developed for
energy production. Switchgrass, miscanthus, and short-rotation tree biomass are
bioenergy feedstocks (Domec et al., 2017). Organic molecules from aquatic plants
and algae form aquatic biomass. These by-products can come from aquaculture,
algae or kelp farming, or algal biofuel production (Sharma et al., 2020a, b).
As the population grows, the demand for food and other resources also increases,
leading to an increase in agricultural and industrial activities. Agricultural wastes
are increasing because they are abandoned, consumed, and handled inefficiently
and infrequently. In poor countries, biomass waste is left in fields to decompose or
burn in the open, harming the environment. The rapid growth of cities worldwide,
and the increasing demand for building materials require more sustainable sources
of energy and raw “materials” (Tripathi et al., 2020). Industrialization, overpopula-
tion, and unhealthy competition harm the environment. These processes produce
many phenols, azo dyes, PAHs, polychlorinated chemicals, insecticides, heavy met-
als, and others (Paital, 2020). These toxins quickly and steadily kill the species that
sustain the ecosystem without degrading. These pollutants have teratogenic, carci-
nogenic, mutagenic, and toxic effects worldwide. This underscores the ecological
treatment of organic wastes (Devi & Kumar, 2020). Chemical and UV degradation,
high-temperature incineration, and pit burial are common methods of xenobiotic
disposal. These technologies have significant drawbacks, including sludge disposal
and harmful by-products. Therefore, we need to quickly and completely replace
harmful physical and chemical processes with ecological ones. This page presents
the enzymes of numerous microorganisms that biodegrade various pollutants, their
uses, and approaches to overcome their limitations (Singh, 2023).

2 Biomass Waste and Its Potential

Biomass waste, organic matter derived from a variety of sources including agricul-
tural residues, forestry by-products, food waste, and municipal solid waste, has
become a growing environmental problem. The sheer volume of biomass waste
Application of Enzymes in Biomass Waste Management 191

generated worldwide poses a major challenge for proper management and disposal.
However, these wastes also represent an untapped resource with enormous potential
for a sustainable and circular economy (Dogu et al., 2021; Karić et al., 2022). It can
be converted into bioenergy through various processes such as anaerobic digestion,
incineration, or gasification. The energy derived from biomass waste can be used for
heat generation, electricity generation, or as biofuel for transportation. Using bio-
mass waste as an energy source can help reduce greenhouse gas emissions, keep
waste out of landfills, and promote sustainable energy production (Huang
et al., 2023).
Composting and fertilizer production are alternative uses for biomass waste
besides energy use. Composting biomass waste produces nutrient-rich soil amend-
ments that increase soil fertility and agricultural productivity. Nitrogen, phospho-
rus, and potassium are extracted from biomass waste and compacted into fertilizers
for crop growth (Mehta et al., 2015; Khan et al., 2023). Bio-based products can also
be produced from biomass waste. Biofuels, bioplastics, and biochemicals are
extracted from biomass waste. Bioethanol and biodiesel from biomass waste can
replace fossil fuels in transportation, reducing greenhouse gas emissions and depen-
dence on nonrenewable resources (Okolie et al., 2021). They could promote soil
fertility, energy production, and bio-based products. By using biomass waste sus-
tainably and efficiently, we can reduce waste, promote circular economy concepts,
and create a more sustainable and resource-efficient future (Chojnacka et al., 2020).
In recent years, the growing awareness of environmental issues and the need for
sustainable practices have led to the development of innovative solutions for bio-
mass waste management. Despite the tremendous potential of biomass waste, there
are several challenges that stand in the way of widespread application. Technological
barriers, economic viability, and the logistics of collection, sorting, and processing
remain the biggest challenges (Yadav et al., 2022). Biomass waste is a valuable
resource that, if managed sustainably, can solve many environmental problems
while contributing to a greener and more sustainable future. By adopting innovative
technologies and promoting circular economy practices, society can unlock the full
potential of biomass waste, reduce its environmental impact, and leverage its value
as a renewable resource. Collaboration between governments, industry, and the
public is essential to foster a paradigm shift toward a more resource-efficient and
environmentally conscious society (Fatimah et al., 2020).

3 Significance of Enzymes in Biowaste Management

The enzyme is an efficient biocatalyst capable of biodegrading substances under


moderate conditions. The active sites of the enzymes are highly selective so that
only certain substrates can be bound, lowering the activation energy required for
catalysis. As a result, the reaction kinetics and specificity of such processes are very
high (Shi et al., 2022). Enzymes can accelerate cellular activities by shortening the
time required to transport substrates within the cell. As the focus has shifted in
recent years to the development of cyclic technologies, the importance of enzymes
192 P. Ranjan et al.

and their applications has increased significantly. For this plan to work, both the
substrate and the end product must be biodegradable and compatible with the con-
cepts of recycling and reuse. Enzymes are biological molecules that, when used
commercially, offer solutions to a wide range of problems (El-Gendi et al., 2021).
Safe disposal of agricultural residues and conversion from synthetic to natural mate-
rials are environmentally friendly. Despite increasing interest in using agricultural
residues in biorefineries, high lignin concentrations and expensive chemical treat-
ment have prevented their use. Biological delignification with xylanase, cellulose,
and ligninolytic enzymes is efficient, cost-effective, and carbon-neutral (Nahak
et al., 2022; Hans et al., 2019). Protein engineering has enabled the synthesis of
industrial enzymes with improved production yields and resistance to harsh condi-
tions that have multiple applications. Enzymatic reactions occur faster and consume
less water and energy than chemical reactions; therefore, enzymes can be reused. In
biological processes, enzymes do not harm other microbes, such as bacteria.
Therefore, enzymatic processes are promising for the treatment of wastewater,
including oils, fats, and organic micropollutants (Guldhe et al., 2015).

4 Potential Microbial Enzymes

Potential microbial enzymes have applications in industry and environmental pro-


tection. Cellulases, amylases, proteases, and lipases are needed for the production
of biofuels, starch, detergents, and biodiesel. Pectinases, xylanases, and ligninases
are used in fruit juice extraction, paper production, and the degradation of lignocel-
lulosic biomass. In agriculture, chitinases can be used as natural pesticides and anti-
fungals. Feed additives such as phytoses increase nutrient availability (Arya et al.,
2022). Anaerobic digestion and biogas production require methane-forming
enzymes. These enzymes are versatile and enable sustainable waste management,
bioremediation, and environmental solutions. As biotechnology improves, new
microbial enzymes are being discovered, increasing their potential for biotechno-
logical, industrial, and environmental applications and creating a more sustainable
future. These enzymes enable optimization of resources, reduction in waste, and
promotion of green technologies in various industries (Chukwuma et al., 2020).
Few of the numerous bacteria and fungi decompose biological waste. These bac-
teria are able to efficiently degrade floral components such as cellulose, hemicellu-
lose, lignin, proteins, and lipids with their enzymes. Biomass waste can be degraded
by bacteria and fungi Cellulomonas, Bacillus, Pseudomonas, Trichoderma,
Aspergillus, and Penicillium. These bacteria thrive under conditions such as aera-
tion, moisture, and temperature that promote growth and enzymatic activity. These
effective decomposers can break down biomass waste, return nutrients to the soil,
and contribute to more sustainable waste management in natural ecosystems or
well-managed composting systems (Table 1). Understanding and utilizing these
specialized microbes can improve biomass waste management and reduce the envi-
ronmental impact of discarded flowers. Although microbes can withstand harsh
Application of Enzymes in Biomass Waste Management 193

Table 1 List of potential microbes and the enzymes secreted by them for the degradation of
biomass waste
Enzymes Microbial sources Role against biomass waste References
Amylases Bacteria: Bacillus subtilis, B. Amylases find use in the food Far et al. (2020)
cereus, B. amyloliquefaciens, industry in the creation of and Kieliszek
B. coagulans, B. polymyxa, B. sugars like glucose and et al. (2020)
acidocaldarius, B. maltose from starch. Laundry
lcheniformis, Lactobacillus detergents employ them to
micrococcus, Pseudomonas, eliminate starch-based stains.
Arthrobacter, Escherichia, They eliminate starch stains in
Proteus, Thermomonospora, detergents. Using yeast
Serretia biomass for waste
Fungi: Aspergillus, management has produced
Penicillium, Cephalosporium, novel metabolites such as
Neurospora, and Rhizopus glucans, minerals,
carotenoids, and enzymes
with several commercial uses
Cellulases Bacteria: Trichoderma viride, This enzyme is very Sampathkumar
Trichoderma reesei, promising for bioconversion et al. (2019) and
Cellulomonas, Cellvibrio, of cellulosic biomass into Rabby et al.
Pseudomonas sp., Bacillus biofuels and bio-based (2022)
sphaericus, Bacillus subtilis, products. They are essential
and Micrococcus for the conversion of
Fungus: Rhizopus oryzae, cellulose-­rich waste into
Aspergillus, Penicillium, biofuels, and bio-based
Cellulomonas, Chaetomium, products have great potential.
Clostridium, They are crucial for
Thermomonospora, bioethanol synthesis by
Trichoderma, Fusarium, and converting cellulose into
Alternaria fermentable sugars. Microbial
cellulases are used in pulp and
paper, textiles, laundry,
biofuels, food and feed,
brewing, and agriculture
Chitinases Bacteria: Actinobacteria like In agriculture, chitinases Joe and Sarojini
Streptomyces aureofaciens, destroy chitin in the cell walls (2017) and Saini
Streptomyces coelicolor, of insects and fungi, acting as et al. (2020)
Streptomyces glaucescens, natural pesticides or
Streptomyces kanamyceticus, antifungals. Chitin is broken
Streptomyces lividans, down into oligomers and
Streptomyces parvies, monomers. They are essential
Streptomyces venezuelae and for the biological conversion
Streptomyces viridifaciens, of chitin waste into
Vibrio harveyi biofertilizers. Chitin and its
Fungus: Trichoderma, derivatives are used in waste
Talaromyces flavus management, wound healing,
biocontrol agents, drug
delivery, and food production,
including sweeteners, growth
factors, and feed additives
(continued)
194 P. Ranjan et al.

Table 1 (continued)
Enzymes Microbial sources Role against biomass waste References
Lipases Bacteria: Bacillus Lipases have application Bashiri et al.
thermocatenulatus, possibilities in various fields, (2022) and Gao
Enterobacter cloacae, such as manufacturing et al. (2020)
detergents, biofuels, and
Acidithiobacillus ferrooxidans,
Roseobacter spp., Bacillus processed foods. They
amyloliquefaciens, accelerate the hydrolysis of
Anoxybacillus gonensis UF7 oils and fats into their
Fungi: A. niger, Burkholderiacomponent fatty acids and
cepacia, Candida rugosa, glycerol. These reactors digest
Chromobacterium viscosum, glycerol and fatty acids to
Geotrichum candidum, Mucor make biodiesel. Fungi are one
javanicus, P. restrictum, P. of the best lipase sources due
camemberti U-159, P. to their substrate selectivity
simplicissimum, Pseudozyma and stability under several
antarctica, Rhizopus delemar,chemical and physical
conditions. Lipase-forming
R. homothallicus, R.s niveus, R.
oligosporus, R. oryzae, R. fungus species cleanse oil and
rhizopodiformis, and dairy products and
Rhizomucor pusillus bioremediation. Filamentous
fungi perform better in
solid-state fermenters than in
submerged ones
Ligninases Bacteria: Clostridium Ligninases degrade lignin, a Nargotra et al.
clariflavum DSM 19732, C. complex polymer found in (2022) and
thermocellum, Microbacterium plant cell walls. This helps in Rodriguez et al.
phyllosphaerae, Rhodococcus bioremediation and (2023)
jostii RHA1, Streptomyces degradation of lignocellulosic
ipomoeae CECT 3341 biomass. The enzymes are
Fungi: A. flavus PN3, used in several industrial
Saccharomyces cerevisiae bioprocesses. In
lignocellulosic biorefineries,
not many biofuels and other
bioproducts can be produced
with microbial enzymes due
to some fundamental
problems. Pretreatment is
often considered the most
expensive step in converting
lignocellulosic biomass into
fermentable sugars. Lignin on
polysaccharides hinders
cellulolytic enzymes. This is
the main reason why enzymes
cannot degrade lignocellulosic
biomass
(continued)
Application of Enzymes in Biomass Waste Management 195

Table 1 (continued)
Enzymes Microbial sources Role against biomass waste References
Methane-­ Bacteria: Methanogenic These enzymes are involved Quezada-­
producing bacteria, genera of Bacteroides, in anaerobic digestion Morales et al.
enzymes Ruminiclostridium, processes and find application (2023) and
Enterococcus, and in biogas production from Wongfaed et al.
Parabacteroides organic waste. One of the (2023)
Fungi: Pleurotus djamor, most promising substrates for
Trichoderma longibrachiatum, biogas production is
and A. niger lignocellulosic biomass.
However, its recalcitrant
structure limits its
effectiveness as a converter.
The purpose of this research is
to create a microbial
consortium that can generate
cellulolytic enzymes while
also investigating the genetic
and taxonomic facets of
lignocellulose degradation
Pectinases Bacteria: Bacillus These enzymes are essential Ibrahim and Ma
licheniformis, B. subtilis, B. in fruit juice extraction, as (2017) and Haile
firmus, Torulopsis kefyr, they help in breaking down and Ayele (2022)
Candida pseudotropicalis var, pectin, a complex
lactosa, and C. pseudotropicalis polysaccharide present in the
Fungi: A.us awamori, A. cell walls of fruits
fumigates, A. giganteus, A. They are gaining a lot of
niger, A. tubingensis, Lentinus interest in the industrial
edodes, Trichoderma reesei, sector, particularly in
Wickerhamomyces anomalous applications such as the textile
Yeast: Saccharomyces fragilis, and fruit processing sectors,
S. thermantitonum as well as oil extraction and
the fermentation of coffee and
tea
Phytases Bacteria: Bacillus subtilis, Phytases are used as feed Vashishth et al.
Escherichia coli, Streptomyces additives because they break (2023)
lividans, Lactobacillus down phytic acid in plant
plantarum feeds, improving the
Fungi: A. niger, P. citrinum, availability of nutrients to
Trichoderma reesei, Rhizopus animals. These enzymes play
oryzae, Pichia pastoris a critical role in catalyzing the
Yeast: S. cerevisiae degradation of phytate
complexes, facilitating the
recycling of phosphate into an
available form in the
ecosystem
(continued)
196 P. Ranjan et al.

Table 1 (continued)
Enzymes Microbial sources Role against biomass waste References
Proteases Bacteria: B. subtilis, E. coli, Proteases are valuable in the Sharma et al.
Streptococcus thermophilus, detergent industry, where they (2020a, b) and
Lactobacillus spp. aid in removing protein-based Solanki et al.
Fungi: Aspergillus spp., stains from fabrics. They also (2021)
Trichoderma spp., Penicillium find applications in the food
spp., Candida spp. industry for tenderizing meat
Yeast: S. cerevisiae, and in pharmaceutical
Kluyveromyces lactis, Pichia research. Leather, textiles,
pastoris detergent, management of
Algae: Chlorella vulgaris, waste, farming, animal
Spirulina platensis husbandry, beauty products,
and pharmaceuticals use
proteases. These enzymes are
economically and
environmentally viable
Xylanases Bacteria: B. subtilis, Xylanases have applications Bhardwaj,
Cellulomonas flavigena, in the pulp and paper industry, Kumar & Verma
Paenibacillus polymyxa, as they break down (2019)
Streptomyces sp. hemicellulose in wood,
Fungi: As. niger, A. oryzae, improving the quality of paper
Penicillium oxalicum, and reducing chemical usage
Thermomyces lanuginosus,
Trichoderma reesei, T.
longibrachiatum

conditions, most of them prefer optimal conditions that are difficult to achieve out-
side the laboratory. Most bioremediation systems are aerobic, but microbial degra-
dation of refractory chemicals can also occur anaerobically. Refractory,
lignin-containing, and organic pollutants are degraded by bacteria and fungi using
various internal and extracellular enzymes (Kumar 2021; Narayanan et al., 2023).
Complex substances are readily degraded and metabolized by these enzymes. Fungi
release extracellular enzymes into the environment, while bacteria use intracellular
enzymes to degrade contaminants. The combined action of these enzymes breaks
down pollutants into less hazardous or nontoxic forms (Narayanan et al., 2023).
This enzymatic remediation method demonstrates how microorganisms can clean
up the environment and improve ecosystems.

5 Basic Thermostable Enzymes Against Biomass Waste

In the biotechnology industry, enzymes are an important tool to address the growing
problem of biomass waste. Biomass wastes, such as agricultural residues and plant
by-products, are abundant but difficult to handle. Basic enzymes offer a promising
solution to this environmental problem. Thermal stability and activity of enzymes at
Application of Enzymes in Biomass Waste Management 197

elevated temperatures are critical properties, especially in processes involving the


conversion of plant biomass into biofuels (Gupta & Verma, 2015). This is because
many of the enzymatic reactions required for biomass degradation and biofuel pro-
duction are carried out at elevated temperatures, a process known as thermophilic or
high-temperature enzymatic hydrolysis. Researchers and biotechnologists are
always looking for and making thermostable enzymes from different sources, such
as thermophilic microorganisms and extremophiles, to make biofuel production and
other biotechnological processes more efficient and sustainable (Sharma et al.,
2019). Here are some key applications of thermostable enzymes in this process:

5.1 High-Temperature Enzymatic Hydrolysis

This process uses thermostable enzymes to break down biomass at high tempera-
tures. Due to increased molecular mobility, enzymatic processes tend to run faster
at higher temperatures. This ensures the rapid degradation of complex polysaccha-
rides such as cellulose and hemicellulose during biomass degradation. Enzymatic
access to biomass can be challenging due to its complex and densely packed struc-
tures, especially for lignocellulosic feedstocks. It can improve the degradation pro-
cess by making substrates more soluble and accessible, allowing enzymes to
penetrate and interact more deeply with the biomass. Many microorganisms cannot
survive under these conditions, so contamination is minimized during enzymatic
hydrolysis. This is useful wherever sterility must be maintained, but especially in
large-scale processes (Kumar & Akhtar, 2019; Bhandari et al., 2021). To streamline
and improve the efficiency of biorefinery processes, hydrolysis can be combined
with other high-temperature processes such as fermentation or downstream process-
ing. High-temperature enzymatic hydrolysis can accelerate reactions and shorten
processing times, saving costs in industrial applications and making biomass-based
processes such as biofuel production more profitable (Verma et al., 2014).

5.2 Lignocellulose Degradation

Lignocellulosic biomass, which consists of cellulose, hemicellulose, and lignin, is


an important feedstock for biofuel production. Lignocellulose degradation is a key
process in biomass conversion and biorefinery applications. Lignocellulose is the
main component of plant cell walls and consists of three main polymers: cellulose,
hemicellulose, and lignin (Okolie et al., 2021; Hemati et al., 2022). These polymers
are tightly interconnected, making the degradation of lignocellulose a complex and
challenging task. Thermostable enzymes such as ligninases help degrade lignin,
which is a barrier to enzymatic access, improving the overall efficiency of biomass
degradation. By catalyzing the degradation of lignin, these enzymes improve the
accessibility of cellulose and hemicellulose to other enzymes, increasing the overall
198 P. Ranjan et al.

efficiency of biomass degradation. The enzymes exhibit remarkable stability and


activity at elevated temperatures, making them invaluable tools in biotechnological
applications (Sindhu et al., 2016).

5.3 Enhanced Substrate Accessibility

The higher temperatures favor the action of thermostable enzymes in biomass deg-
radation. At higher temperatures, the solubility of the biomass substrates increases,
making them more accessible to the enzymes (Baramee et al., 2020). This makes it
easier for thermostable enzymes to get into the complex structures of biomass and
interact with the target parts, such as cellulose and hemicellulose. As a result, enzy-
matic hydrolysis is accelerated, leading to a higher yield of fermentable sugars. This
increased efficiency is critical for biofuel production and other biotechnological
processes, as it improves the overall productivity and cost-effectiveness of convert-
ing biomass into valuable products. The improved enzymatic activity at higher tem-
peratures leads to faster degradation of biomass components such as cellulose,
hemicellulose, and lignin. This leads to higher production rates of fermentable sug-
ars, which are important precursors for biofuels and biochemicals (Saravanan et al.,
2022). Faster enzymatic hydrolysis enables more efficient use of feedstocks and
shorter processing times, contributing to higher overall productivity.

5.4 Reduced Risk of Contamination

Enzymes derived from extremely thermophilic bacteria have long been of techno-
logical interest because of their ability to catalyze industrially important processes
at high temperatures. Enzymatic processes at high temperatures are less susceptible
to contamination by competing microorganisms due to extreme conditions. The use
of thermostable enzymes reduces the risk of undesirable microbial growth and
ensures a more controlled and sterile environment (Chen & Jiang, 2018). Unlike
microbial processes, which can be susceptible to contamination, enzymes act as
biocatalysts, and their activity is tightly controlled. In addition, when thermostable
enzymes are used at elevated temperatures, the risk of contamination is further min-
imized, as many microorganisms cannot survive or thrive under such extreme con-
ditions. This reduced risk of contamination during enzymatic biomass degradation
provides a more controlled and sterile environment, resulting in higher product
yields, increased process efficiency, and improved overall biotechnological applica-
tions such as biofuel production and biorefinery processes. It also promotes sustain-
ability by avoiding the use of harsh chemicals or the generation of toxic by-products,
contributing to a cleaner and more environmentally friendly approach to biomass
utilization (Usmani et al., 2021).
Application of Enzymes in Biomass Waste Management 199

5.5 Process Efficiency and Cost-Effectiveness

Thermostable enzymes enable the enzymatic degradation of biomass at higher tem-


peratures, resulting in faster reaction rates and shorter processing times. They
remain active and stable at high temperatures, allowing them to catalyze reactions
faster. The increased molecular motion and improved enzyme-substrate interactions
at elevated temperatures result in more efficient biomass degradation, leading to
higher productivity and faster conversion of complex biomass components
(Saravanan et al., 2022). The use of thermostable enzymes in biomass degradation
reduces the time required for enzymatic hydrolysis. Shorter processing times lead to
faster production cycles, which can be beneficial on an industrial scale, resulting in
cost savings and improved overall process efficiency. This increased efficiency
leads to cost savings and improved overall economics of the biofuel production
process (Neumann et al., 2022).

5.6 Process Integration

Thermostable enzymes can be used in high-temperature processes, making it easy


to combine many phases in biorefinery ideas in an efficient, resource-efficient, and
environmentally friendly manner (Hemati et al., 2022). This integration maximizes
biomass conversion, waste reduction, biofuel production, and bioproduct value,
making it more sustainable and environmentally friendly. Thermostable enzymes
are seamlessly integrated into biorefinery steps to rapidly and thoroughly convert
biomass into valuable products. This maximizes biomass utilization and minimizes
feedstock waste. Thermostable enzymes efficiently break down biomass pieces in
biorefineries and reduce waste. By converting more biomass into valuable products,
the amount of waste is reduced, making the business more sustainable and resource-
efficient (Saini et al., 2022).

6 Common Biomass Waste Management Practices

Common biomass waste management practices include sustainable and environ-


mentally friendly approaches to efficiently treating organic waste. Composting is a
widely used method that converts biomass waste into nutrient-rich organic fertilizer
through natural decomposition. Anaerobic digestion is another process that decom-
poses biodegradable waste and produces biogas for renewable energy. Biomass-to-­
energy processes, such as pyrolysis and gasification, convert waste into biochar and
syngas for power generation. In addition, biofuel production uses biomass waste as
feedstock for bioethanol and biodiesel. These processes promote the principles of
the circular economy, reduce greenhouse gas emissions, and contribute to a more
sustainable and resource-efficient waste management system (Fig. 1).
200 P. Ranjan et al.

Fig. 1 Different significant biomass waste management practices

Waste Prevention and Minimization The first step in biomass waste management
is to reduce the amount of waste. This can be achieved through efficient agricultural
and forestry practices, and by promoting waste reduction and recycling at the
source, for example, by encouraging composting of food waste. Adopting sustain-
able waste management practices, such as optimizing harvesting techniques and
applying circular economy principles, helps minimize biomass waste, conserve
resources, and reduce environmental impacts. By prioritizing waste prevention and
reduction, we are paving the way for a more sustainable and responsible approach
to biomass waste management.

Biomass Recycling Various forms of biomass waste offer opportunities for recy-
cling or reuse. Agricultural residues, such as crop stalks and husks, find use as ani-
mal feed, bedding material, or in the manufacture of bio-based products. Food waste
can be composted to produce nutrient-rich soil improvers, contributing to sustain-
able waste management and resource utilization (Campos et al., 2019). These recy-
cling practices minimize waste, conserve resources, and support a circular economy
approach that promotes environmental sustainability.

Anaerobic Digestion Anaerobic digestion is a biological process that breaks down


organic materials in the absence of oxygen, producing biogas (primarily methane)
and nutrient-rich digestate. Biomass waste such as food waste or manure can be
anaerobically digested to produce renewable energy and valuable soil amendments.
This sustainable process not only helps manage waste and reduce greenhouse gas
Application of Enzymes in Biomass Waste Management 201

emissions but also harnesses the potential of biomass waste to produce biogas, a
clean and renewable energy source that contributes to a greener and more resource-­
efficient future (Kerkel et al., 2021).

Incineration and Cogeneration Biomass waste can be burned under controlled


conditions to generate heat and electricity. This process, known as biomass combus-
tion or waste to energy, helps reduce dependence on fossil fuels while providing
efficient waste management. Combined heat and power (CHP) further maximizes
benefits by using waste heat during combustion for additional power generation or
industrial applications (Adrio & Demain, 2014). Incineration and cogeneration
offer an environmentally friendly approach to biomass waste management, contrib-
uting to sustainable energy production and reducing greenhouse gas emissions,
making them valuable components of a circular and resource-efficient waste man-
agement system.

Biochemical Conversion Biomass waste can be converted into biofuels and bio-
chemicals through processes such as fermentation or hydrolysis. Lignocellulosic
biomass, including crop residues or dedicated energy crops, can be converted to
bioethanol or biobutanol through fermentation. These bioconversion processes pro-
vide sustainable alternatives to fossil fuels, reduce greenhouse gas emissions, and
promote a greener energy landscape.

Gasification Biomass waste can be subjected to gasification, a thermal process that


converts solid biomass into a synthetic gas or “syngas” consisting of carbon monox-
ide and hydrogen, and other gases. Syngas can be used for the production of elec-
tricity, heat, or biofuels. Gasification offers a versatile and efficient method to utilize
biomass waste, providing an alternative to fossil fuels and reducing greenhouse gas
emissions. Moreover, the process can accommodate a wide range of feedstocks,
making it suitable for diverse biomass sources and contributing to a more sustain-
able and renewable energy future.

Landfilling If other waste management options are not available or feasible, bio-
mass waste may be disposed of in landfills. However, this is generally considered
the least preferred option due to the potential for greenhouse gas emissions, odors,
and contamination of soil and water. As biomass waste decomposes anaerobically
in landfills, it produces methane, a potent greenhouse gas with a significant impact
on climate change. To minimize the environmental impact of landfilling, waste
authorities and governments must prioritize waste prevention, recycling, and sus-
tainable alternatives to reduce the amount of biomass waste sent to landfills.
It is important to note that the choice of biomass waste management method
depends on various factors such as the type and quantity of waste, local regulations,
infrastructure availability, and environmental considerations. The aim is to priori-
tize practices that minimize waste, maximize resource recovery, and minimize envi-
ronmental impacts.
202 P. Ranjan et al.

7 Conclusion

The importance of enzymes and their applications has increased significantly over
the last decade, and now researchers are turning their attention to developing closed-­
loop technologies. For this concept to work, it is important to ensure that both the
starting material and the final product are biodegradable and compatible with the
idea of recycling and reuse. Enzymes are biological molecules that, when used com-
mercially, have the potential to solve a variety of problems. Some examples of these
problems include the disposal of agricultural residues and the substitution of syn-
thetic processes with natural, more environmentally friendly processes. Despite the
growing interest, the high lignin content of the materials and the high cost of chemi-
cal treatment have hindered the use of agricultural residues in biorefineries. The use
of xylanase, a substance called ligninolytic enzymes, in biological delignification is
not only efficient but also cost-effective and carbon-neutral. Protein engineering, for
example, has made it possible to synthetically produce enzymes that can be used
industrially, have a higher production yield, and survive in harsh environments;
these enzymes have a wide range of applications.

References

Adrio, J. L., & Demain, A. L. (2014). Microbial enzymes: Tools for biotechnological processes.
Biomolecules, 4(1), 117–139.
Antero, R. V. P., Alves, A. C. F., de Oliveira, S. B., Ojala, S. A., & Brum, S. S. (2020). Challenges
and alternatives for the adequacy of hydrothermal carbonization of lignocellulosic biomass in
cleaner production systems: A review. Journal of Cleaner Production, 252, 119899.
Arya, P. S., Yagnik, S. M., Rajput, K. N., Panchal, R. R., & Raval, V. H. (2022). Valorization of
agro-food wastes: Ease of concomitant-enzymes production with application in food and bio-
fuel industries. Bioresource Technology, 361, 127738.
Baramee, S., Siriatcharanon, A. K., Ketbot, P., Teeravivattanakit, T., Waeonukul, R., Pason, P.,
et al. (2020). Biological pretreatment of rice straw with cellulase-free xylanolytic enzyme-­
producing Bacillus firmus K-1: Structural modification and biomass digestibility. Renewable
Energy, 160, 555–563.
Bashiri, R., Allen, B., Shamurad, B., Pabst, M., Curtis, T. P., & Ofiţeru, I. D. (2022). Looking for
lipases and lipolytic organisms in low-temperature anaerobic reactors treating domestic waste-
water. Water Research, 212, 118115.
Bhandari, S., Poudel, D. K., Marahatha, R., Dawadi, S., Khadayat, K., Phuyal, S., et al. (2021).
Microbial enzymes used in bioremediation. Journal of Chemistry, 2021, 1–7.
Bhardwaj, N., Kumar, B., & Verma, P. (2019). A detailed overview of xylanases: An emerging
biomolecule for current and future prospective. Bioresources and Bioprocessing, 6(1), 1–36.
Campos, K. R., Coleman, P. J., Alvarez, J. C., Dreher, S. D., Garbaccio, R. M., Terrett, N. K.,
et al. (2019). The importance of synthetic chemistry in the pharmaceutical industry. Science,
363(6424), eaat0805.
Chen, G. Q., & Jiang, X. R. (2018). Next generation industrial biotechnology based on extremo-
philic bacteria. Current Opinion in Biotechnology, 50, 94–100.
Chojnacka, K., Moustakas, K., & Witek-Krowiak, A. (2020). Bio-based fertilizers: A practical
approach towards circular economy. Bioresource Technology, 295, 122223.
Application of Enzymes in Biomass Waste Management 203

Chukwuma, O. B., Rafatullah, M., Tajarudin, H. A., & Ismail, N. (2020). Lignocellulolytic
enzymes in biotechnological and industrial processes: A review. Sustainability, 12(18), 7282.
Devi, P., & Kumar, P. (2020). Concept and application of phytoremediation in the fight of heavy
metal toxicity. Journal of Pharmaceutical Sciences and Research, 12(6), 795–804.
Dogu, O., Pelucchi, M., Van de Vijver, R., Van Steenberge, P. H., D’hooge, D. R., Cuoci, A., et al.
(2021). The chemistry of chemical recycling of solid plastic waste via pyrolysis and gasifica-
tion: State-of-the-art, challenges, and future directions. Progress in Energy and Combustion
Science, 84, 100901.
Domec, J. C., Ashley, E., Fischer, M., Noormets, A., Boone, J., Williamson, J. C., et al. (2017).
Productivity, biomass partitioning, and energy yield of low-input short-rotation American syca-
more (Platanus occidentalis L.) grown on marginal land: Effects of planting density and simu-
lated drought. Bioenergy Research, 10, 903–914.
El-Gendi, H., Saleh, A. K., Badierah, R., Redwan, E. M., El-Maradny, Y. A., & El-Fakharany,
E. M. (2021). A comprehensive insight into fungal enzymes: Structure, classification, and their
role in mankind’s challenges. Journal of Fungi, 8(1), 23.
Far, B. E., Ahmadi, Y., Khosroshahi, A. Y., & Dilmaghani, A. (2020). Microbial alpha-amylase pro-
duction: Progress, challenges and perspectives. Advanced Pharmaceutical Bulletin, 10(3), 350.
Fatimah, Y. A., Govindan, K., Murniningsih, R., & Setiawan, A. (2020). Industry 4.0 based sus-
tainable circular economy approach for smart waste management system to achieve sustainable
development goals: A case study of Indonesia. Journal of Cleaner Production, 269, 122263.
Gao, X. Y., Liu, X. J., Fu, C. A., Gu, X. F., Lin, J. Q., Liu, X. M., et al. (2020). Novel strategy
for improvement of the bioleaching efficiency of Acidithiobacillus ferrooxidans based on the
AfeI/R quorum sensing system. Minerals, 10(3), 222.
Guldhe, A., Singh, B., Mutanda, T., Permaul, K., & Bux, F. (2015). Advances in synthesis of
biodiesel via enzyme catalysis: Novel and sustainable approaches. Renewable and Sustainable
Energy Reviews, 41, 1447–1464.
Gupta, A., & Verma, J. P. (2015). Sustainable bio-ethanol production from agro-residues: A review.
Renewable and Sustainable Energy Reviews, 41, 550–567.
Haile, S., & Ayele, A. (2022). Pectinase from microorganisms and its industrial applications. The
Scientific World Journal, 2022, 1881305.
Hans, M., Kumar, S., Chandel, A. K., & Polikarpov, I. (2019). A review on bioprocessing of paddy
straw to ethanol using simultaneous saccharification and fermentation. Process Biochemistry,
85, 125–134.
Hemati, A., Nazari, M., Asgari Lajayer, B., Smith, D. L., & Astatkie, T. (2022). Lignocellulosics in
plant cell wall and their potential biological degradation. Folia Microbiologica, 67(5), 671–681.
Huang, Y., Shi, Y., & Xu, J. (2023). Integrated district electricity system with anaerobic diges-
tion and gasification for bioenergy production optimization and carbon reduction. Sustainable
Energy Technologies and Assessments, 55, 102890.
Ibrahim, N. E., & Ma, K. (2017). Industrial applications of thermostable enzymes from extremo-
philic microorganisms. Current Biochemical Engineering, 4(2), 75–98.
Joe, S., & Sarojini, S. (2017). An efficient method of production of colloidal chitin for enumeration
of chitinase producing bacteria. Journal of Sciences, 4(16), 37–45.
Karić, N., Maia, A. S., Teodorović, A., Atanasova, N., Langergraber, G., Crini, G., et al. (2022).
Bio-waste valorisation: Agricultural wastes as biosorbents for removal of (in) organic pollut-
ants in wastewater treatment. Chemical Engineering Journal Advances, 9, 100239.
Karlsson, I., Rootzén, J., & Johnsson, F. (2020). Reaching net-zero carbon emissions in con-
struction supply chains – Analysis of a Swedish road construction project. Renewable and
Sustainable Energy Reviews, 120, 109651.
Kerkel, F., Markiewicz, M., Stolte, S., Müller, E., & Kunz, W. (2021). The green platform mol-
ecule gamma-valerolactone–ecotoxicity, biodegradability, solvent properties, and potential
applications. Green Chemistry, 23(8), 2962–2976.
Khan, N., Bolan, N., Jospeh, S., Anh, M. T. L., Meier, S., Kookana, R., et al. (2023). Complementing
compost with biochar for agriculture, soil remediation and climate mitigation. Advances in
Agronomy, 179, 1–90.
204 P. Ranjan et al.

Kieliszek, M., Piwowarek, K., Kot, A. M., & Pobiega, K. (2020). The aspects of microbial biomass
use in the utilization of selected waste from the agro-food industry. Open Life Sciences, 15(1),
787–796.
Kumar, M. (2021). In silico efficacy of [S]-8-gingerol (a derivative) with 6-gingerol against
PT-domain of Polyketide synthase A (PksA). IP International Journal of Medical Microbiology
and Tropical Diseases, 7, 62–64.
Kumar, M., & Akhtar, M. S. (2019). Application of microbial biotechnology in improving salt
stress and crop productivity. Salt Stress, Microbes, and Plant Interactions: Mechanisms and
Molecular Approaches, 2, 133–159.
Mehta, C. M., Khunjar, W. O., Nguyen, V., Tait, S., & Batstone, D. J. (2015). Technologies to
recover nutrients from waste streams: A critical review. Critical Reviews in Environmental
Science and Technology, 45(4), 385–427.
Nahak, B. K., Preetam, S., Sharma, D., Shukla, S. K., Syväjärvi, M., Toncu, D. C., et al. (2022).
Advancements in net-zero pertinency of lignocellulosic biomass for climate neutral energy
production. Renewable and Sustainable Energy Reviews, 161, 112393.
Narayanan, M., Ali, S. S., & El-Sheekh, M. (2023). A comprehensive review on the potential of
microbial enzymes in multipollutant bioremediation: Mechanisms, challenges, and future pros-
pects. Journal of Environmental Management, 334, 117532.
Nargotra, P., Sharma, V., Lee, Y. C., Tsai, Y. H., Liu, Y. C., Shieh, C. J., et al. (2022). Microbial
Lignocellulolytic enzymes for the effective valorization of lignocellulosic biomass: A review.
Catalysts, 13(1), 83.
Neumann, J., Petranikova, M., Meeus, M., Gamarra, J. D., Younesi, R., Winter, M., et al. (2022).
Recycling of lithium-ion batteries – Current state of the art, circular economy, and next genera-
tion recycling. Advanced Energy Materials, 12(17), 2102917.
Okolie, J. A., Nanda, S., Dalai, A. K., & Kozinski, J. A. (2021). Chemistry and specialty industrial
applications of lignocellulosic biomass. Waste and Biomass Valorization, 12, 2145–2169.
Paital, B. (2020). Nurture to nature via COVID-19, a self-regenerating environmental strategy of
environment in global context. Science of the Total Environment, 729, 139088.
Quezada-Morales, D. L., Campos-Guillén, J., De Moure-Flores, F. J., Amaro-Reyes, A., Martínez-­
Martínez, J. H., Chaparro-Sánchez, R., et al. (2023). Effect of pretreatments on the production
of biogas from Castor waste by anaerobic digestion. Fermentation, 9(4), 399.
Rabby, M. R. I., Ahmed, Z. B., Paul, G. K., Chowdhury, N. N., Akter, F., Razu, M. H., et al. (2022).
A combined study on optimization, in silico modeling, and genetic modification of large scale
microbial cellulase production. Biochemistry Research International, 2022, 4598937.
Rasul, G., & Sharma, B. (2016). The nexus approach to water–energy–food security: An option for
adaptation to climate change. Climate Policy, 16(6), 682–702.
Rodriguez, A., Hirakawa, M. P., Geiselman, G. M., Tran-Gyamfi, M. B., Light, Y. K., George,
A., et al. (2023). Prospects for utilizing microbial consortia for lignin conversion. Frontiers in
Chemical Engineering, 5, 1086881.
Saini, S., Chand, M., Sharma, H. O., & Kumar, P. (2020). Role of Chitinases as a waste man-
agement to control global crisis. International Journal for Environmental Rehabilitation and
Conservation, XI, 303–313.
Saini, J. K., Kaur, A., & Mathur, A. (2022). Strategies to enhance enzymatic hydrolysis of lignocel-
lulosic biomass for biorefinery applications: A review. Bioresource Technology, 360, 127517.
Sampathkumar, K., Kumar, V., Sivamani, S., & Sivakumar, N. (2019). An insight into fungal cel-
lulases and their industrial applications. In Approaches to enhance industrial production of
fungal cellulases (pp. 19–35). Springer.
Saravanan, A., Kumar, P. S., Jeevanantham, S., Karishma, S., & Vo, D. V. N. (2022). Recent
advances and sustainable development of biofuels production from lignocellulosic biomass.
Bioresource Technology, 344, 126203.
Secondi, L., Principato, L., Ruini, L., & Guidi, M. (2019). Reusing food waste in food manufactur-
ing companies: The case of the tomato-sauce supply chain. Sustainability, 11(7), 2154.
Sharma, S., Vaid, S., Bhat, B., Singh, S., & Bajaj, B. K. (2019). Thermostable enzymes for indus-
trial biotechnology. In Advances in enzyme technology (pp. 469–495). Elsevier.
Application of Enzymes in Biomass Waste Management 205

Sharma, B., Brandt, C., McCullough-Amal, D., Langholtz, M., & Webb, E. (2020a). Assessment
of the feedstock supply for siting single-and multiple-feedstock biorefineries in the USA and
identification of prevalent feedstocks. Biofuels, Bioproducts and Biorefining, 14(3), 578–593.
Sharma, A., Jamali, H., Vaishnav, A., Giri, B. S., & Srivastava, A. K. (2020b). Microbial biofilm:
An advanced eco-friendly approach for bioremediation. In New and future developments in
microbial biotechnology and bioengineering: Microbial biofilms (pp. 205–219). Elsevier.
Shi, Y., Zhou, Y., Lou, Y., Chen, Z., Xiong, H., & Zhu, Y. (2022). Homogeneity of supported
single-­atom active sites boosting the selective catalytic transformations. Advanced Science,
9(24), 2201520.
Sindhu, R., Binod, P., & Pandey, A. (2016). Biological pretreatment of lignocellulosic biomass –
An overview. Bioresource Technology, 199, 76–82.
Singh, D. (2023). Advances in industrial waste management. In Waste management and resource
recycling in the developing world (pp. 385–416). Elsevier.
Soffian, M. S., Halim, F. Z. A., Aziz, F., Rahman, M. A., Amin, M. A. M., & Chee, D. N. A. (2022).
Carbon-based material derived from biomass waste for wastewater treatment. Environmental
Advances, 9, 100259.
Solanki, P., Putatunda, C., Kumar, A., Bhatia, R., & Walia, A. (2021). Microbial proteases:
Ubiquitous enzymes with innumerable uses. 3 Biotech, 11(10), 428.
Tripathi, S., Lohia, P. & Dwivedi, D.K., (2020). Contribution to sustainable and environmental
friendly non-toxic CZTS solar cell with an innovative hybrid buffer layer. Solar Energy, 204,
748–760.
Usmani, Z., Sharma, M., Awasthi, A. K., Sivakumar, N., Lukk, T., Pecoraro, L., et al. (2021).
Bioprocessing of waste biomass for sustainable product development and minimizing environ-
mental impact. Bioresource Technology, 322, 124548.
Vashishth, A., Tehri, N., Tehri, P., Sharma, A., Sharma, A. K., & Kumar, V. (2023). Unravelling the
potential of bacterial phytases for sustainable management of phosphorous. Biotechnology and
Applied Biochemistry, 70(5), 1690–1706.
Verma, J. P., Jaiswal, D. K., & Sagar, R. (2014). Pesticide relevance and their microbial degrada-
tion: A-state-of-art. Reviews in Environmental Science and Bio/Technology, 13, 429–466.
Wongfaed, N., O-Thong, S., Sittijunda, S., & Reungsang, A. (2023). Taxonomic and enzymatic
basis of the cellulolytic microbial consortium KKU-MC1 and its application in enhancing bio-
methane production. Scientific Reports, 13(1), 2968.
Yadav, S., Patel, S., Killedar, D.J., Kumar, S. & Kumar, R., (2022). Eco-innovations and sustain-
ability in solid waste management: An indian upfront in technological, organizational, start-ups
and financial framework. Journal of Environmental Management, 302, 113953.
Zyryanov, M., Shvetsov, V., Milyaeva, I., & Dozhdev, E. (2021). Improving logging process effi-
ciency in the context of rational natural resource management. In E3S web of conferences (Vol.
285, p. 06002). EDP Sciences.
Pretreatment Techniques for Derivation
of Value-Added Products from Agro-Waste
Biomass

Tran Thi Viet Ha, Nguyen Thi An Hang, and Nguyen Minh Viet

1 Introduction

Agricultural wastes are the by-products of growing and processing crops, animals,
and other raw agricultural products such as fruits and vegetables. Agricultural
wastes can be found in large quantities. They are the results of the production and
processing of agricultural products, but they are not regarded as products in and of
themselves. These outputs may contain substances that are useful to humans; how-
ever, their monetary values fall short of the expenses incurred in the process of
gathering, transporting, and processing these compounds for useful applications.
Their consistency can be liquids, slurries, or solids, depending on the system and the
kind of agricultural operations that are being carried out, and their composition will
be determined by those factors. Agricultural waste, also known as agro-waste, con-
sists of discarded animal products (such as manure and animal carcasses), waste
generated during the processing of food (just 20% of maize is canned, while 80% of
it is squandered), waste from agricultural production (such as maize stalks, bagasse
from sugarcane, drops and culls from fruits and vegetables, and pruning), and haz-
ardous agricultural waste (including, but not limited to, herbicides, pesticides, and
insecticides among other things). Although estimates of the amount of garbage pro-
duced by agriculture are difficult to come by, it is widely accepted that this type of
waste accounts for a sizeable fraction of the overall waste matter in developed

T. T. Viet Ha (*) · N. Thi An Hang


Faculty of Advanced Technology and Engineering, VNU Vietnam Japan University,
Hanoi, Vietnam
e-mail: ttv.ha@vju.ac.vn
N. Minh Viet
VNU Key Laboratory of Advanced Material for Green Growth, VNU University of Science,
Hanoi, Vietnam

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 207
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_11
208 T. T. Viet Ha et al.

countries. Growing agricultural production has unavoidably led to a growth in the


quantities of waste produced by livestock, as well as by agricultural crop wastes and
by-products of agro-industrial processes.
Biomass refers to the total quantity of organic matter derived from any form of
biological material that can be utilized for the production of bioproducts.
Lignocellulose constitutes the primary structural constituent of arboreal vegeta-
tion. Non-lignified plants, including grass, serve as a significant and renewable
source of organic matter. Lignocellulose is composed of lignin, hemicellulose,
and cellulose, rendering it a highly valuable substrate in the field of biotechnol-
ogy. Rice straw is a plentiful lignocellulosic waste material globally, derived as a
by-product of rice cultivation, and possesses significant potential as a
bioresource.
Currently, the topic of waste conversion and utilization is not novel. Efficient
management and utilization of agricultural wastes hold promise as a sustainable
resource for value addition, making significant contributions to energy supply secu-
rity and ecological sustainability. The revitalization of the economy, the protection
of natural resources, the expansion of the circular economy, and the improvement of
people’s physical and mental well-being are all potential outcomes of the conver-
sion and use of agricultural wastes in an acceptable manner. Lignocellulosics are
regarded as recalcitrant compounds that exhibit resistance to degradation due to
their intricate chemical composition. The prevention of effective utilization of cel-
lulose for ethanol production is attributed to the presence of a crystalline structure
consisting of cellulose fibrils surrounded by hemicellulose, as well as the presence
of the lignin seal. These factors impede the penetration of degrading enzymes,
thereby hindering the process.
This chapter aims to present the various methods and technologies that may be
used to pretreat agricultural waste in order to facilitate the process of converting
this waste into a variety of goods. The main aim of pretreating agricultural wastes
and other lignocellulosic biomass is to attain various advantages, such as reducing
energy consumption during the conversion procedure, cost mitigation, and pro-
moting direct sugar production from the biomass. The implementation of pretreat-
ment procedures is crucial in guaranteeing the protection of the sugar generated
from degradation while also minimizing the production of substances that hinder
the process. Various pretreatment methods are commonly employed, encompass-
ing physical, chemical, biological, physicochemical, and green solvent-based pro-
cesses. The findings of this chapter will make a valuable contribution to the
advancement of academic research by offering current insights into the diverse
pretreatment approaches employed for agricultural waste. The findings will also
serve as a catalyst for further exploration into the efficacy of waste conversion as
a means of waste management. Farmers, environmentalists, legislators, and small
company owners, among others, will find this study’s conclusions useful in their
respective fields.
Pretreatment Techniques for Derivation of Value-Added Products from Agro-Waste… 209

2 Pretreatment Techniques for Agriculture Waste

The pretreatment stage holds significant importance in agricultural biomass thermo-


chemical and biochemical conversion. This stage primarily involves making struc-
tural modifications that specifically target the inherent resistance of biomass
materials. The implementation of retreatment procedures is imperative to optimize
biomass qualities and properties, thereby maximizing its efficacy in energy utiliza-
tion. There exist multiple approaches to pretreatment processes (Fig. 1).

3 Physical Pretreatment

Physical methods are recognized as having the greatest potential among all pretreat-
ments for transforming food debris and lignocellulosic residues. The utilization of
physical pretreatment methods circumvents the need for chemical agents, thereby
mitigating the production of waste materials and inhibitors that could impede sub-
sequent reactions. Mechanical, microwave, ultrasound, and thermal pretreatments
are widely employed techniques to enhance the efficacy of primary stages in bio-
mass processing.
Typically, the use of raw biomass for energy generation is hindered by several
factors, including its elevated moisture content, limited bulk density, diminished
energy content, uneven combustion characteristics, and particulate matter emis-
sions. The challenges mentioned above pose difficulties in the efficient use of raw

Fig. 1 Diverse pretreatment techniques for the conversion of agro-waste biomass


210 T. T. Viet Ha et al.

biomass, specifically in terms of handling, storing, and transportation. These chal-


lenges can be effectively addressed through the process of densification (Singh
et al., 2023), cutting, milling, chopping, shearing, and stirring (Dahunsi, 2019).
For example, mechanical milling considerably increases the digestibility of pre-
treated materials, making it one of the many technologies that have already been
studied and shown to be useful as biomass pretreatments. In order to alter the physi-
cal structure of biomass, reduce particle size, loosen material structure, increase
specific surface area, and improve enzyme accessibility, this technique is crucial (Ji
et al., 2017). Another example is about microwave processing. Due to the irradiation
uniformly heating the materials, a loose lignocellulosic structure can be formed
(Chen & Wan, 2018). Microwaves are a form of non-ionizing radiation character-
ized by wavelengths ranging from 0.01 to 1 meter and frequencies between 300 and
300,000 megahertz. Despite using minimal pretreatment severities, this method
improves hydrothermal biomass fractionation by encouraging interaction between
substrate and reaction media (Tsubaki et al., 2016). In the presence of ions or polar
molecules, microwave heating supports two major mechanisms that generate fast
heating: ionic conduction and dipole rotation. The microwave system gains from the
novel heating mechanism’s decreased heating time, efficiency in volumetric heating
that is both uniform and selective, and increased effectiveness in transferring energy.
Hydrothermal pretreatment methods are also considered a popular and effective
physical method, the temperature for heating is approximately 120–180 °C, heating
for a certain time depends on the biomass, and the water molecules through ligno-
cellulose run to promote its hydrolysis (Yu et al., 2022).

4 Chemical Pretreatment

Chemical pretreatment is conducted to separate the distinct biopolymeric composi-


tion of the biomass material. Due to their widespread availability, high treatment
capacity, and cost-effectiveness, chemicals such as acids, alkalis, ionic liquids, and
organic solvents are commonly used in chemical pretreatment procedures. Even low
concentrations of diluted acids are adequate to achieve a significant level of efficacy
in treatment, and complete efficiency is attained using this approach.

4.1 Acid Pretreatment

Various acids have been used as reagents to degrade lignocellulose by dissolving


preferentially hemicellulose. The primary parameters employed for pretreatment
included acid concentration ranging from 1% to 10%, treatment time duration span-
ning from 5 to 120 min, temperature ranging from 80 to 190 °C, and low-speed
stirring. An elevated concentration level results in the phenomenon of inhibition. In
most cases, HCl and H2SO4 in dilute concentrations are used in the acid treatment
Pretreatment Techniques for Derivation of Value-Added Products from Agro-Waste… 211

process (Binod et al., 2010; Chen et al., 2011). In some situations, organic acids
oxalic, citric, and acetic acid have also been shown to be helpful (Peng et al., 2019).
The advantages of this method include the efficient removal of hemicelluloses,
the enhancement of biomass porosity, the generation of a high glucose yield, and the
elimination of the requirement for a separate enzymatic hydrolysis step, as this
treatment also performs hydrolysis. Nevertheless, this particular approach appears
to be less efficient when applied to lignin. Additionally, it necessitates the use of
corrosion-resistant equipment, which incurs higher costs. The recovery of acid in
this process is also expensive, and there is a greater likelihood of the formation of
inhibitory furfural. Furthermore, the chemicals employed in this method are costly.

4.2 Alkali Pretreatment

The process entails subjecting the material to a diluted alkaline solution, such as
NaOH, KOH, or Ca(OH)2, at a moderate temperature (Shetty, 2017; Cheng et al.,
2010; Gundupalli, 2021). This treatment effectively eliminates a significant portion
of the lignin and partially removes hemicellulose. The parameters for pretreatments
involve the use of a low concentration of alkali, preferably ranging from 1% to 10%.
The duration of the pretreatment can vary from 5 min to 5 days, depending on the
specific requirements. It is recommended to maintain a moderate temperature dur-
ing the process and in the condition of stirring slowly. It should be noted that a sig-
nificant concentration of alkali has the potential to impede the efficacy of the
pretreatment procedure.
The utilization of this particular method necessitates the implementation of
milder conditions in comparison to acidic pretreatment, while also facilitating a
greater degree of lignin removal. However, sometimes the catalysts needed for the
process are quite expensive.

4.3 Oxidative Pretreatment

The process of ozonation applied to biomass represents a sustainable approach for


enhancing the enzymatic hydrolysis of lignocellulosic materials. Ozone, a potent
oxidizing agent with an oxidation potential of 2.07 V at 25 °C, can be readily gener-
ated from oxygen. Ozone exhibits a high degree of selectivity when interacting with
various functional groups, displaying a preference for attacking phenols such as
lignin rather than carbohydrates. Its affinity for phenols is measured at 1300 M−1
s−1, whereas its affinity for sucrose and ethanol is measured at 0.5 and 0.45 M−1 s−1,
respectively. The degradation of phenols in aqueous solution through ozonation is
characterized by three distinct phases. The initial phase, which lasts several min-
utes, is marked by rapid phenol degradation. The presence of hydroxyl radical
(•OH) in significant quantities is observed as a result of the ozonolysis of phenols in
212 T. T. Viet Ha et al.

an aqueous solution. It is worth noting that the pH level significantly influences both
the specificity of ozonation and the decomposition of ozone. Therefore, the pres-
ence of an acidic solution leads to a targeted ozonation process, while a basic solu-
tion results in a nonspecific reaction with hydroxyl radicals (•OH) (Mvula & Von
Sonntag, 2003; Brolin et al., 1993). Consequently, it is recommended to incorporate
a buffer for pH regulation in the ozonation process.
In some other research studies, the oxidative pretreatment technique used the
oxidizing characteristics of hydrogen peroxide (H2O2) and ozone gas to break down
the lignin and hemicellulose present, thereby promoting the liberation of chemicals
that can dissolve. H2O2 undergoes a chemical decomposition process, resulting in
the formation of hydroxide ions (•OH) and superoxide ions (•O2-). This decomposi-
tion process aids in the breakdown of lignin structure while concurrently putting an
end to the generation of any by-products that have an inhibiting effect.

5 Biological Pretreatment

The biological pretreatment method entails the application of enzymes, which are
synthesized by microorganisms, specifically fungi, to break down, depolymerize,
and cleave various biomass constituents, including cellulose, hemicellulose, and
lignin. The aforementioned method of pretreatment offers several advantages in
comparison to the other two methods mentioned. These advantages include the
absence of toxic by-products, enhanced productivity of desired products and
reduced energy consumption, substrate specificity, and an efficient reaction process.
However, there are several limitations that restrict the effectiveness of biological
pretreatment processes, including the slow nature of the process, the need for sig-
nificant space, and the requirement to maintain optimal conditions for microbial
growth. Typically, the residence time for this process is approximately 14 days. In
order to address the consumption of organic content in biomass by microbes, it is
necessary to consider the biological pretreatment process from additional perspec-
tives, such as techno-economic considerations. Therefore, this methodology is not
frequently employed by researchers and is considered less appealing in comparison
to alternative approaches.

5.1 Fungal Pretreatment

Fungi are widely recognized microorganisms due to their ability to interact with and
degrade lignocellulosic residue through the action of different enzymes. These
fungi are frequently encountered in natural environments, with many of them capa-
ble of producing a variety of cellulolytic, hemicellulolytic, and ligninolytic enzymes.
The lignocellulolytic fungi include species from the ascomycetes (e.g., Aspergillus
sp., Penicillium sp., and Trichoderma reesei), basidiomycetes including white-rot
Pretreatment Techniques for Derivation of Value-Added Products from Agro-Waste… 213

fungi (e.g., Schizophyllum sp. and P. chrysosporium), brown-rot fungi (e.g.,


Fomitopsis palustris), and few anaerobic species (e.g., Orpinomyces sp.) (Dashtban
et al., 2009; Paudel & Qin, 2015). Furthermore, the use of fungal pretreatment in
conjunction with moderate pretreatments such as alkaline or ethanol-based organo-
solv pretreatment resulted in comparable sugar yields to those obtained from ther-
mochemical pretreatments, approximately ranging from 90% to 100%. This implies
that the utilization of various pretreatment methods, along with the incorporation of
mediators or other chemical agents to augment ligninolytic activity, has the poten-
tial to enhance the efficiency of fungal pretreatment, making it economically viable.
The primary concern remains the duration of this pretreatment method, which has
not undergone substantial reduction in the majority of recent scholarly
investigations.
The fungal pretreatment method is also possible to be conducted in both solid or
liquid form, with the former being the preferred method due to its ability to replicate
the natural environment, accommodate higher substrate loading, and avoid the gen-
eration of liquid waste streams. The growth of fungi and the depolymerization of
lignin are influenced by various factors, including moisture content, temperature,
and aeration. The size of biomass particles also plays a significant role in the pro-
cess, primarily due to their impact on the surface area available for exposure. An
appropriate particle size ensures optimal exposure without obstructing the airflow,
which is necessary due to the extremely aerobic nature of the oxidation process.
Prior to fungal inoculation, decontamination of the feedstock is typically required
due to the susceptibility of white-rot fungi to being easily outcompeted by other
microorganisms present in the biomass.

5.2 Bacteria Pretreatment

Certain bacterial species, predominantly found in soil environments, possess the


capability to enzymatically degrade lignin. These fungi exhibit a higher growth rate
compared to the majority of other fungi and possess the ability to break down lignin
into smaller fragments that are soluble in water. These fragments can be extracted
and utilized as valuable products, allowing for the creation of enzymes that can be
used effectively within the biorefinery framework.
Sphingobium sp. SYK-6, an α-proteobacteria, is widely recognized as one of the
most extensively studied bacterial strains with well-documented capabilities in lig-
nin depolymerization, and extensive research has been conducted on the degrada-
tion pathways of biaryls and monoaryls (87). However, the understanding of
bacterial ligninolytic systems remains limited (Bugg et al., 2011). Additional bacte-
rial species, namely, Pseudomonas putida mt-2 and Rhodococcus jostii RHA1, have
been recognized for their aptitude for lignin degradation within lignocellulosic bio-
mass. These strains do not require hydrogen peroxide to depolymerize lignin, indi-
cating that their ligninolytic system likely consists of laccases or other nonperoxidases
(Ahmad et al., 2010).
214 T. T. Viet Ha et al.

5.3 Microbial Consortium Pretreatment

Due to the intricate and diverse composition of lignin, microorganisms have evolved
a considerably nonspecific mechanism for its degradation, relying on extracellular
enzymes to facilitate oxidative reactions. The phenomenon under discussion is
commonly known as enzymatic combustion. Unlike typical enzymatic reactions
that exhibit specificity and follow a well-defined metabolic pathway, enzymatic
combustion encompasses a collection of nonspecific reactions that give rise to a
multitude of intermediate products. The specific composition of these intermediates
varies considerably depending on factors such as the substrate involved, the strain
of the enzyme, and the prevailing environmental conditions.
Microbial depolymerization is a biodegradation process that can be facilitated by
various microorganisms, including fungi and bacteria. However, it is primarily
wood decay fungi that play a significant role in this particular process. The majority
of these organisms belong to the taxonomic group Basidiomycetes, which exhibit
the ability to overcome challenges in their growth, such as limited nitrogen avail-
ability and the existence of toxic substances (Martínez, 2005). White-rot fungi have
been identified as the most efficient microorganisms for lignin degradation
(Reid, 1995).
The significance of microbial degradation of lignin in the biogeochemical carbon
cycle stems from the fact that lignin constitutes approximately one-third of the over-
all carbon assimilated through photosynthesis. Furthermore, the degradation of lig-
nin enables the hydrolysis of various carbon sources, such as cellulose. The use of
microbial depolymerization has a lot of promise for usage in a range of lignocellu-
losic biomass-using businesses, including but not limited to the production of paper
and biofuels (Martínez et al., 2009).

5.4 Enzyme Pretreatment

Rather than cultivating microorganisms within the biomass, an alternative approach


involves incorporating enzymatic extracts into the feedstock to facilitate the depo-
lymerization of lignin. This approach has the potential to decrease the duration of
pretreatment and minimize carbohydrate degradation, thereby streamlining the pro-
cessing procedure (Chen et al., 2010).
The efficacy of this technology is typically inferior to that of microbial pretreat-
ment due to challenges related to heterogeneous catalysis and the intricate nature of
the ligninolytic enzyme system. The latter typically necessitates the cooperative
action of multiple enzymes and the presence of mediators to achieve efficient lignin
decomposition. Efficient production of enzymes is also necessary for enzymatic
pretreatment. Nevertheless, the synthesis of these enzymes typically relies on the
consumption of specific nutrients, such as nitrogen, and is associated with second-
ary metabolism rather than primary growth. Consequently, the majority of
Pretreatment Techniques for Derivation of Value-Added Products from Agro-Waste… 215

microorganisms that possess the ability to degrade lignin inherently exhibit limited
production of ligninolytic enzymes, resulting in inconsistent yields attributed to
intricate regulatory mechanisms. Furthermore, the retrieval of ligninolytic enzymes
poses a significant challenge despite their extracellular nature.

6 Physicochemical Pretreatment

6.1 Steam Explosion Pretreatment

In this method, the feedstock undergoes a treatment involving the application of


high saturated steam pressure ranging from 0.7 to 4.8 MPa, along with a high tem-
perature ranging from 160 to 260 °C, for a duration of a few seconds to a few min-
utes. Subsequently, the pressure is rapidly released as part of this process. The
composition of biomass endures disintegration as a result of the significant mechan-
ical and chemical changes that occur during this particular process, leading to the
occurrence of hydrolysis and subsequent liberation of hemicellulose. This technique
is particularly well suited for hardwood materials; however, it should be noted that
it may produce inhibitors when exposed to elevated temperatures (Barbanera
et al., 2015).

6.2 Alkali-Heat Pretreatment

The parameters for the pretreatment process include a range of concentrations for
ammonia, varying from 15% to 40%. The temperature during pretreatment can be
adjusted within the range of 40–180 °C, while the pressure can be set between 0.1
and 2.2 MPa. The duration of the pretreatment process typically lasts between 30
and 60 min. Ammonia has been found to be efficacious in the treatment of lignocel-
lulosic biomass. There exist three distinct methodologies for employing ammonia-­
based treatment, namely (i) ammonia fiber expansion (AFEX), (ii) ammonia recycle
percolation (ARP), and (iii) soaking aqueous ammonia (SAA) (Naik, 2021).

6.3 Extrusion Pretreatment

The process of extrusion involves the integration of multiple operations within a


single unit. The inlet of the extruder is where unprocessed materials, such as food
for animals, are loaded. These materials are subsequently conveyed through the bar-
rel by a feeding screw mechanism. As the material traverses the barrel, it experi-
ences the effects of frictional heat, blending, and shearing. The procedure can be
216 T. T. Viet Ha et al.

modified by altering the screw’s design. The center of the barrel is designated as the
compression zone, while the end is designated as the expansion/wearing zone. The
aforementioned phenomenon leads to significant abrasion of the material upon the
release of pressure, thereby causing the degradation of its structural components.
The depolymerization of cellulose, hemicellulose, and protein and lignin occurs
during the extrusion process (Camire, 1998). Furthermore, the potential thermal
degradation of sugars and amino acids may occur, depending on the magnitude of
the stresses exerted.

7 Green Solvent-Based Pretreatment

7.1 Ionic Liquid Pretreatment

The technique of ionic liquid engineering employs ion solvents with lower melting
points and non-volatile properties, which offer advantages for lignocellulosic com-
ponents in hydrolysis reactions. The parameters for pretreatments include a recom-
mended concentration range of 5–10% of solid loading for the ionic liquid. The
treatment time should be within the range of 1–5 h, while maintaining a temperature
of 120 °C. It is important to ensure that the melting point of the ionic liquid is below
100 °C. Additionally, it has been noted that giving the mixture a little stir now and
again has a beneficial effect. The aforementioned methods of treatment are consid-
ered significantly more effective compared to the previous chemical methods, par-
ticularly for lignocellulosic waste materials such as wheat, corn, and wood. The
process of biomass gasification using materials such as straw, switchgrass, miscan-
thus, and other types of leaves has significantly increased after pretreatment with
1-N-butyl-imidazolium chloride. The higher costs associated with such a technique
that solely utilizes solvents have been addressed by incorporating ammonia as a
complementary component (Nguyen et al., 2010). The majority of these techniques
have been employed for ethanol processing, and thus, their efficacy in biomass
remains unexplored. It is predicted that there is maximum potential for it to be rec-
ommended for the utilization of lignocellulosic feedstock in agricultural animal
feed due to its sorting capability (Hartmann et al., 2000).

7.2 Deep Eutectic Solvent

Abbott et al. introduced deep eutectic solvents in 2003, precisely aligning with the
12 principles of green chemistry. These solvents have been recognized as a novel
class of environmentally friendly solvents due to their desirable characteristics such
as low vapor pressure, cost-effectiveness, facile synthesis, and biodegradability
(Abbott et al., 2004; Tang et al., 2017). In recent years, the utilization of dilute acid
Pretreatment Techniques for Derivation of Value-Added Products from Agro-Waste… 217

pretreatment has gained significant popularity in the field of biomass deconstruc-


tion, particularly for the purpose of liberating fermentable sugars that can be subse-
quently utilized in bioenergy production.
Deep eutectic solvents exhibit physical and chemical properties that bear resem-
blance to, and are commonly regarded as, a novel iteration of ionic liquids. Moreover,
deep eutectic solvents present themselves as an environmentally sustainable substi-
tute for traditional ionic liquids. Deep eutectic solvents consist of a combination of
hydrogen bond donors, such as urea, amides, alcohols, acids, and others, along with
hydrogen bond acceptors, such as choline chloride, betaine, alanine, and others.
This method has garnered a great deal of attention due to its exceptional ability to
solubilize lignin from LCB, low cost of synthesis, stability, satisfactory recyclabil-
ity, and eco-friendly properties. The utilization of deep eutectic solvents for the
extraction of polysaccharides and lignin from biomass holds promise for the pro-
duction of biofuels and bioproducts that possess significant commercial value
(Sharma et al., 2022).

7.3 Supercritical Fluid

Supercritical fluids refer to substances that exist in a state above their critical tem-
perature and pressure thresholds. Under these specified conditions, the fluid in ques-
tion does not exhibit a vapor-liquid phase transition. Instead, it exists solely in a
homogeneous phase state, wherein its properties, including diffusivity, viscosity,
and density, fall within the intermediate range between those of gases and liquids.
Supercritical fluids can be modified in terms of their properties through the manipu-
lation of pressure or temperature. Additionally, the introduction of specific liquid
solvents can further influence these properties by exploiting the chemical interac-
tions between the modifiers and the solutes. The most prevalent supercritical fluids
include carbon dioxide, ammonia, water, and hydrocarbons such as propane and
butane (Walsh et al., 1987). Among them, carbon dioxide is the most used super-
critical fluid due to its non-toxic, recyclable, low-cost, environmentally friendly
characteristics (considering it in a closed loop as a solvent), and low critical tem-
perature and pressure (31.1 °C and 7.36 MPa, respectively) (Daza Serna, 2016).

8 Conclusion and Future Perspectives

In general, the preparatory procedure can be categorized as shown in Table 1.


Agriculture and forestry are significant contributors to environmental waste,
making them potential sources for bioenergy production. These resources have been
employed in developing nations following the implementation of pretreatment tech-
niques to optimize the efficient utilization of biomass. Numerous comprehensive
elucidations regarding pretreatment techniques and diverse characterization
218 T. T. Viet Ha et al.

Table 1 Strategies for the pretreatment of agro-waste biomass


Pretreatment methods
Green
Physical Chemical Biological Physicochemical solvent-based
Objectives Decrease the Hydrolyze Degrade lignin Breakdown Breakdown
size of lignin, from lignin-­ lignin-­
particles, hemicellulose, holocellulose holocellulose holocellulose
enhance the and cellulose components linkages linkages
surface area,
and diminish
the crystallinity
of cellulose
Examples Mechanical Alkaline Fungal Steam explosion Green
Microwave Acid Termites Alkali heat solvent-based
Ultrasound Oxidative Microbial Extrusion pretreatment
Thermal Organic consortium …. Supercritical
…. solvent Enzymes fluid
…. ….. Deep eutectic
solvent
Ionic liquid
pretreatment
…..
Pros Environmental Less Environmental Less Higher
friendly hazardous friendly corrosiveness efficiency
The efficacy is procedure No chemical Higher energy Short process
notably high Less requirement efficiency time
No chemical by-product Low energy Short process
necessity degradation consumption time
Rapid
procedure
duration
High selectivity
and uniformity
Cons High energy Toxic Long time Chemical Difficult to
requirement Corrosiveness Large space recovery control and
Expensive of apparatus requirement High operation handle
Taking a long Partial cost Expensive
time hydrolysis of Formation of
Recovery of hemicellulose inhibitors
chemicals Continuous
monitoring of
microbial
growth is
required
Potential
health risks

processes, focusing on the efficacy of lignin removal, have been extensively docu-
mented. The pretreatment procedure plays a crucial role in promoting the biode-
gradability of biomass, increasing biofuel production, and minimizing the overall
processing duration. The utilization of a composite approach involving multiple
Pretreatment Techniques for Derivation of Value-Added Products from Agro-Waste… 219

techniques has been observed to yield greater efficacy in enhancing the digestibility
of lignocellulosic biomass. There are several different factors that can be used to
evaluate the effectiveness of pretreatment, which may differ across different meth-
ods and are contingent upon the specific characteristics of the biomass being treated.
The efficacy of the pretreatment process is contingent upon the specific characteris-
tics of the biomass being utilized. The identification of a commercially practicable
method poses a significant challenge for the community of scientists. The discovery
of a straightforward, cost-effective, highly productive, and reliable pretreatment
technique would have a significant impact on the development of renewable energy,
specifically in the field of biofuel production. Furthermore, the pursuit of this objec-
tive aligns with the Sustainable Development Goals (SDGs) established by the
United Nations, specifically Goal 7, which aims to promote affordable and clean
energy. Further investigation is necessary in order to identify an appropriate and
effective pretreatment technique for lignocellulosic biomass, as it plays a crucial
role in augmenting biogas production. The achievement of commercial viability
will remain an elusive goal unless the pretreatment method demonstrates promise.
In addition to the existing laboratory-scale research, there is a requirement for pilot-
level work, which is a really challenge to the scientist and engineering society.

References

Abbott, A. P., Boothby, D., Capper, G., Davies, D. L., & Rasheed, R. K. (2004). Deep eutectic
solvents formed between choline chloride and carboxylic acids: Versatile alternatives to ionic
liquids. Journal of the American Chemical Society, 126, 9142–9147.
Ahmad, M., Taylor, C. R., Pink, D., Burton, K., Eastwood, D., Bending, G. D., & Bugg,
T. D. H. (2010). Development of novel assays for lignin degradation: Comparative analysis of
bacterial and fungal lignin degraders. Molecular BioSystems, 6, 15–821.
Barbanera, M., Buratti, C., Cotana, F., Foschini, D., & Lascaro, E. (2015). Effect of steam explo-
sion pretreatment on sugar production by enzymatic hydrolysis of olive tree pruning. In 69th
conference of the Italian Thermal Engineering Association, ATI 2014 (Vol. 81, pp. 146–154).
Binod, P., Sindhu, R., Singhania, R. R., Vikram, S., Devi, L., Nagalakshmi, S., Kurien, N.,
Sukumaran, R. K., & Pandey, A. (2010). Bioethanol production from rice straw: An overview.
Bioresource Technology, 101, 4767–4774.
Brolin, A., Gierer, J., & Zhang, O. Y. (1993). On the selectivity of ozone delignification of soft-
wood kraft pulps. Wood Science and Technology, 27, 115–129.
Bugg, T. D. H., Ahmad, M., Hardimana, E. M., & Rahmanpoura, R. (2011). Pathways for degrada-
tion of lignin in bacteria and fungi. Natural Product Reports, 28, 1883–1896.
Camire, M. E. (1998). Chemical changes during extrusion cooking: Recent advances. In Process-­
induced chemical changes in food (pp. 109–121). Springer.
Chen, Z., & Wan, C. (2018). Ultrafast fractionation of lignocellulosic biomass by microwave-­
assisted deep eutectic solvent pretreatment. Bioresource Technology, 250, 532–537.
Chen, S., Zhang, X., Singh, D., Yu, H., & Yang, X. (2010). Biological pretreatment of lignocel-
lulosics: Potential, progress and challenges. Biofuels, 1, 177–199.
Chen, W.-H., Tu, Y.-J., & Sheen, H.-K. (2011). Disruption of sugarcane bagasse lignocellulosic
structure by means of dilute sulfuric acid pretreatment with microwave-assisted heating.
Applied Energy, 88, 2726–2734.
220 T. T. Viet Ha et al.

Cheng, Y.-S., Zheng, Y., Chao Wei, Y., Dooley, T. M., Jenkins, B. M., & Vandergheynst,
J. S. (2010). Evaluation of high solids alkaline pretreatment of rice straw. Applied
Biochemistry and Biotechnology, 162, 1768–1784.
Dahunsi, S. O. (2019). Mechanical pretreatment of lignocelluloses for enhanced biogas produc-
tion: Methane yield. Bioresource Technology, 280, 18–26.
Dashtban, M., Schraft, H., & Qin, W. (2009). Fungal bioconversion of lignocellulosic residues;
opportunities & perspectives. International Journal of Biological Sciences, 5, 578–595.
Daza Serna, L. V., Orrego Alzate, C. E., & Cardona Alzate, C. A. (2016). Supercritical fluids as
a green technology for the pretreatment of lignocellulosic biomass. Bioresource Technology,
199, 113–120.
Gundupalli, M. P., Cheng, Y.-S., Chuetor, S., Bhattacharyya, D., & Sriariyanun, M. (2021). Effect
of dewaxing on saccharification and ethanol production from different lignocellulosic biomass.
Bioresource Technology, 339, 125596.
Hartmann, H., Angelidaki, I., & Ahring, B. K. (2000). Increase of anaerobic degradation of par-
ticulate organic matter in full-scale biogas plants by mechanical maceration. Water Science and
Technology, 41, 145–153.
Ji, G., Han, L., Gao, C., Xiao, W., Zhang, Y., & Cao, Y. (2017). Quantitative approaches for illus-
trating correlations among the mechanical fragmentation scales, crystallinity and enzymatic
hydrolysis glucose yield of rice straw. Bioresource Technology, 241, 262–268.
Martínez, A. T., Speranza, M., Ruiz-Dueñas, F. J., Ferreira, P., Camarero, S., Guillén, F., Martínez,
M. J., Gutiérrez, A., & Del Río, J. C. (2005). Biodegradation of lignocellulosics: Microbial,
chemical, and enzymatic aspects of the fungal attack of lignin. International Microbiology, 8,
195–204.
Martínez, Á. T., Ruiz-Dueñas, F. J., Martínez, M. J., Del Río, J. C., & Gutiérrez, A. (2009).
Enzymatic delignification of plant cell wall: From nature to mill. Current Opinion in
Biotechnology, 20, 348–357.
Mvula, E., & Von Sonntag, C. (2003). Ozonolysis of phenols in aqueous solution. Organic &
Biomolecular Chemistry, 1, 1749–1756.
Naik, G. P., Poonia, A. K., & Chaudhari, P. K. (2021). Pretreatment of lignocellulosic agricultural
waste for delignification, rapid hydrolysis, and enhanced biogas production: A review. Journal
of the Indian Chemical Society, 98, 100147.
Nguyen, T.-A. D., Kim, K.-R., Han, S. J., Cho, H. Y., Kim, J. W., Park, S. M., Park, J. C., & Sim,
S. J. (2010). Pretreatment of rice straw with ammonia and ionic liquid for lignocellulose con-
version to fermentable sugars. Bioresource Technology, 101, 7432–7438.
Paudel, Y. P., & Qin, W. (2015). Two bacillus species isolated from rotting wood samples are
good candidates for the production of bioethanol using agave biomass. Journal of Microbial &
Biochemical Technology, 7, 219–225.
Peng, J., Abomohra, A. E.-F., Elsayed, M., Zhang, X., Fan, Q., & Ai, P. (2019). Compositional
changes of rice straw fibers after pretreatment with diluted acetic acid: Towards enhanced bio-
methane production. Journal of Cleaner Production, 230, 775–782.
Reid, I. D. (1995). Biodegradation of lignin. Canadian Journal of Botany, 73, 1011–1018.
Sharma, V., Tsai, M.-L., Chen, C.-W., Sun, P.-P., Patel, A. K., Singhania, R. R., Nargotra, P., &
Dong, C.-D. (2022). Deep eutectic solvents as promising pretreatment agents for sustainable
lignocellulosic biorefineries: A review. Bioresource Technology, 360, 127631.
Shetty, D. J., Kshirsagar, P., Tapadia-Maheshwari, S., Singh, S. K., & Dhakephalkar, P. K. (2017).
Alkali pretreatment at ambient temperature: A promising method to enhance biomethanation
of rice straw. Bioresource Technology, 226, 80–88.
Singh, R., Kumar, R., Sarangi, P. K., Kovalev, A. A., & Vivekanand, V. (2023). Effect of physical
and thermal pretreatment of lignocellulosic biomass on biohydrogen production by thermo-
chemical route: A critical review. Bioresource Technology, 369, 128458.
Tang, X., Zuo, M., Li, Z., Liu, H., Xiong, C., Zeng, X., Sun, Y., Hu, L., Liu, S., & Lei, T. (2017).
Green processing of lignocellulosic biomass and its derivatives in deep eutectic solvents.
ChemSusChem, 10, 2696–2706.
Pretreatment Techniques for Derivation of Value-Added Products from Agro-Waste… 221

Tsubaki, S., Azuma, J.-I., Yoshimura, T., Maitani, M., Suzuki, E., Fujii, S., & Wada, Y. (2016).
Microwave-induced biomass fractionation. In Biomass fractionation technologies for a ligno-
cellulosic feedstock based biorefinery. Elsevier.
Walsh, J. M., Ikonomou, G. D., & Donohue, M. D. (1987). Supercritical phase behavior: The
entrainer effect. Fluid Phase Equilibria, 33, 295–314.
Yu, C., Li, M., Zhang, B., Xin, Y., Tan, W., Meng, F., Hou, J., & He, X. (2022). Hydrothermal
pretreatment contributes to accelerate maturity during the composting of lignocellulosic solid
wastes. Bioresource Technology, 346, 126587.
Significance of Enzymatic Actions
in Biomass Waste Management:
Challenges and Future Scope

Prangya Rath, Laxmi Kant Bhardwaj , Mini Chaturvedi,


and Abhishek Bhardwaj

1 Introduction

Several anthropogenic activities such as agriculture practices and industrialization


are increasing day by day. These activities contribute to increasing the level of pol-
lutants in the environment. These pollutants may be azo dyes, phenols, polycyclic
aromatic hydrocarbons (PAHs), polychlorinated biphenyls (PCBs), pesticides,
dioxins, and heavy metals (Bhardwaj et al., 2023a). There are several processes
such as physical and chemical treatments and activated carbon adsorption that have
been developed in recent years for removing pollutants from waste. Industrial
wastes contain toxic organic and inorganic materials. The presence of organic mate-
rials in the waste may be treated biologically for the minimization of the overall
treatment cost.
Enzymes can play an important role in the treatment of waste. The use of enzymes
in waste treatment was first proposed in 1930. They are target-specific and can par-
ticularly attack target pollutants. They are used as an efficient catalyst and are used
at a particular pH and temperature. Researchers were focusing on the development
of new technology for waste treatment that will be faster, cheaper, simpler, and

P. Rath
Amity Institute of Environmental Sciences (AIES), Amity University,
Noida, Uttar Pradesh, India
L. K. Bhardwaj (*)
Amity Institute of Environmental, Toxicology, Safety, & Management (AIETSM), Amity
University, Noida, Uttar Pradesh, India
M. Chaturvedi
Department of Hematology, Delhi Paramedical Management Institute (DPMI),
New Delhi, India
A. Bhardwaj
Department of Environmental Science, Amity University, Gwalior, Madhya Pradesh, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 223
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_12
224 P. Rath et al.

more reliable. Now, they have developed enzymatic processes with the help of bio-
technological tools. This technique replaces costly chemicals and does not require
specialized machinery. It is a recommended technique for its clean/green and biode-
gradable nature. It also has several advantages. It falls between the physicochemical
and biological processes because it contains chemical processes that are based on
the biological catalysis action.
Enzymatic conversion of biomass offers several advantages over chemical and
physical techniques. These advantages are as follows:
(i) It can be operated at high and low concentrations of contaminant.
(ii) It could be operated over a broader range of parameters including pH, tempera-
ture, and salinity.
(iii) There are no shock-loading effects.
(iv) There is a reduction in the volume of sludge.
(v) The process of controlling is simple.
(vi) It is an environmentally friendly technique.
With these potential advantages, researchers have focused on the use of this tech-
nique for the treatment of wastewater, solid wastes, and hazardous wastes.
Biomass waste treatment uses enzymes from microorganisms such as bacteria or
fungi. They are involved in several metabolic activities and produce different
enzymes. These enzymes are specific to their work and involve a series of chemical
reactions. Enzymes that are produced by aerobic bacteria (Alcaligenes,
Mycobacterium, Pseudomonas, Rhodococcus, and Sphingomonas) are used in the
degradation of pesticides and hydrocarbons, while enzymes that are produced by
anaerobic bacteria are used in the degradation of compounds such as polychlori-
nated biphenyls (PCBs), dichlorination of trichloroethylene (TCE), and chloroform
(Sharma, 2012).
The pollutants including azo dyes, phenols, PAHs, pesticides, polychlorinated
compounds and heavy metals have been studied to know their teratogenic nature,
carcinogenic potential, and mutagenic, as well as their toxic effects on the health of
humans (Liu et al., 2019; Alam et al., 2023; Bhardwaj et al., 2023b). However, the
enzymatic technique is slow, and only some bacterial species are capable enough to
produce specific enzymes. So, the researchers prefer genetically engineered
microbes for this process. This process converts the pollutants from toxic forms to
nontoxic forms (Phale et al., 2019). Thermophilic bacteria Caldicellulosiruptor
bescii have been studied to directly convert plant biomass into bioethanol. This
could be used as a potential agent in the commercial sector for bioethanol produc-
tion (Chung et al., 2014). Similarly, hemicellulases, cellulases, xylanases, endox-
ylanases, and β-xylosidases extracted from a variety of ascomycetes (Trichoderma,
Aspergillus, etc.) have been applied extensively at commercial levels to act as bio-
catalysts in lignocellulosic biorefineries (Ferreira et al., 2016). A recent study has
shown that the co-cultivation and harvesting of microalgae and filamentous fungi
were efficiently used for wastewater treatment and biofuel production (Chu
et al., 2021).
Significance of Enzymatic Actions in Biomass Waste Management: Challenges… 225

The aim of this chapter is to focus on different types of enzymes and their appli-
cations for waste treatment.

2 Enzymes Used for Biomass Conversion, Degradation,


and Hydrolysis

Enzymes can help in converting biomass into various useful by-products. Biomass
conversion involves the catalysis of complex organic materials, such as plant fibers,
feedstocks, and agricultural waste, into simple components. These components
could further be processed into biofuels, biochemicals, and other valuable products
(Demirbas, 2009; Mahapatra et al., 2021). Several types of enzymes are commonly
used in biomass conversion processes (Table 1). Some of the examples of enzymes
are given below:

Table 1 List of the different enzymes with their sources and applications
S. no. Name of enzyme Sources Application of enzymes References
1 Alkylsulfatase Pseudomonas C12B Surfactant degradation Toesch et al.
(2014)
2 Azoreductase Intestinal microflora Removal of azo dyes Sandhya (2010)
3 α-Amylase Bacillus subtilis Glucose production and Kolusheva and
4 Glucoamylase hydrolysis of starch Marinova
(2007) and
Presečki et al.
(2013)
5 Cellulase Trichoderma Hydrolysis of cellulose in Champagne and
6 Cellobio-­ harzianum, sludges and municipal solid Li (2009), Khan
hydrolase Trichoderma viride waste (MSW) to produce et al. (2016) and
7 Cellobiose alcohol, sugars, and energy Pandey et al.
(2017)
8 Exo-1,4-b-D-­
glucosidase
9 Chitinase Streptomyces Production of N-acetyl Mander et al.
anulatus CS242 glucosamine from shellfish (2016)
waste through bioconversion
10 Chloro-­ Caldariomyces Oxidation of phenolic Sjoblad and
peroxidase fumago compounds Bollag (2021)
11 Cyanidase Pseudomonas sp. Cyanide decay Akcil et al.
(2003)
12 Cyanide Fusarium lateritium Cyanide hydrolysis Ebbs (2004)
hydratase
13 Depolymerase Bacteriophage Bacterial exopolysaccharideKnecht et al.
(2020)
14 Hemoglobin Blood Removal of aromatic amines Pérez-Prior
and phenols et al. (2014)
(continued)
226 P. Rath et al.

Table 1 (continued)
S. no. Name of enzyme Sources Application of enzymes References
15 L-Galactono-­ Candida Conversion of galactose Nicell (2003)
lactone oxidase norvegensis from L-ascorbic acid
16 Laccase Pleurotus Decolorization of kraft, Khatami et al.
(P. ostreatus, removal of phenols, (2022)
P. pulmonarius) bleaching of paper pulp,
and Trametes binding of phenols and
(T. versicolor, aromatic amines with
T. hirsuta) humus, detoxification of
wastewater
17 Lactases Bacterial Dairy waste processing and Coughlin and
production of value-added Charles (2019)
products
18 Lignin peroxidase Phanerochaete Removal of phenols and Falade et al.
(LiP) chrysosporium aromatic compounds, (2017)
decolorization of Kraft
bleaching effluents
19 Lipase Various sources Improved sludge dewatering Nimkande and
20 Lysozyme Bacterial Bafana (2022)
and Dai et al.
(2023)
21 Manganese Phanerochaete Decolorization of synthetic Bansal and
peroxidase (MnP) chrysosporium dyes, removal of phenolic Kanwar (2013)
contaminants, removal of
endocrine disruptive
chemicals (EDC),
degradation of chlorinated
alkanes and alkenes,
degradation of chlorinated
dioxins
22 Nitrile hydratase Mesorhizobium sp. Removal of acrylonitrile Feng et al.
(2008)
23 Parathion Azohydromonas Hydrolyzation of Zhao et al.
hydrolase australica organophosphate pesticides (2021)
24 Pectin lyase Clostridium Pectin degradation Yadav et al.
beijerinckii (2009)
25 Pectin Aspergillus niger Kohli et al.
methylesterase (2015)
26 Phosphatase Escherichia coli Removal of heavy metals Chaudhuri et al.
C90 (2013)
27 Phosphoesterases Aspergillus sydowii Removal of chlorpyrifos, Soares et al.
CBMAI 935 diazinon, parathions (2021)
28 Polyphenol Mushroom Removal of phenolic Li et al. (2021)
oxidase (Agaricus bisporus) compounds
29 Proteases Bacillus Hydrolyze or break down Karn and
licheniformis, the protein molecules in Kumar (2015)
Aspergillus niger, meat, biodegradation of the and Arslan et al.
Chlorella vulgaris industrial sludge (2021)
30 Tyrosinase Mushroom Removal of phenolic Bayramoglu
(Agaricus bisporus) compounds et al. (2013)
Significance of Enzymatic Actions in Biomass Waste Management: Challenges… 227

Amylases They are extracted from fungal and bacterial strains that hydrolyze
starch into glucose and maltose. Starch is often used as a feedstock in biomass con-
version processes, and amylases are employed at a commercial scale to convert it
into fermentable sugars for biofuel/bioethanol production (Castro et al., 2011).

Cellulases They help in the breakdown of cellulose into glucose moieties, and
cellulose is a major component of plant cell walls formed by β-1,4-glycosidic
bonds. They consist of three main types: endoglucanases, exoglucanases (or
cellobiohydrolases), and β-glucosidases (Dashtban et al., 2010). Cellulases
have been observed to be important in biofuel production such as cellulosic
ethanol and in the conversion of biomass into various biochemicals (Siqueira
et al., 2020).

Hemicellulases They target hemicellulose, which is another complex carbohy-


drate present in plant cell walls. These enzymes include xylanases, mannanases, and
arabinases (Zanuso et al., 2021). Bacteria and fungi produce large amounts of hemi-
cellulases. They break down hemicellulose into smaller sugar units, such as xylose,
mannose, and arabinose, which can be fermented into biofuels or used as building
blocks for biochemical production (Méndez-Líter et al., 2021).

Ligninases They include lignin peroxidases and manganese peroxidases and are
involved in the degradation of lignin, which is a complex and highly resistant poly-
mer found in plant biomass. Lignin allows access to the cellulose and hemicellulose
components for further enzymatic degradation or chemical processing (Siqueira
et al., 2020; Bilal & Iqbal, 2021).

Proteases They are also known as peptidases and are responsible for breaking
down proteins into smaller peptides and further to amino acids.

Laccases They are extracellular enzymes containing multi-copper that consist of


glycoproteins with dimeric, tetrameric, and monomeric units characterized by bac-
teria, fungi, and plants (Shekher et al., 2011). Lignin and phenolic compounds,
which are found in banana peels, sawdusts, and rice bran, have enhanced production
of laccase (Muthukumarasamy et al., 2015).

3 Mechanism of Treatment of Biomass

The enzymes break down complex organic compounds into simpler components, as
depicted in Fig. 1. The mechanism involves the following steps:
228 P. Rath et al.

Fig. 1 Diagrammatic representation of the mechanism of biomass waste treatment

3.1 Pretreatment of Biomass Wastes

Biomass wastes such as agricultural residues, wooden chips, and residual crops are
pretreated to enhance the enzyme’s accessibility to the complex carbohydrates pres-
ent in biomass. Such techniques include combinations of physical, chemical, or
biological means, for example, acid or alkali treatment, fungal/bacterial/yeast treat-
ment, milling, and steam explosion. These methods are capable of disrupting the
biomass structure, thereby making it more susceptible to enzymatic degradations
(Zhang et al., 2021).

3.2 Enzyme Production

Microorganisms that are capable of producing enzymes suitable for biomass degra-
dation are cultivated in a controlled environment. These microbial cultures are
grown in bioreactors under specific conditions that promote enzyme production
either naturally or by genetic engineering (Guo et al., 2023).

3.3 Enzymatic Hydrolysis

The pretreated biomass and enzyme solutions are mixed to initiate the hydrolysis
process. Enzymes, such as cellulases, hemicellulases, and lignin, act on complex
carbohydrates of biomass and hydrolyze them into simpler molecules. These
enzymes work synergistically to degrade different components of the biomass,
releasing glucose, xylose, and other sugars (Huang et al., 2022a).
Significance of Enzymatic Actions in Biomass Waste Management: Challenges… 229

3.4 Fermentation and Further Processing

The resultant sugar-rich hydrolysate is subjected to fermentation using microorgan-


isms, such as yeast or bacteria, to convert the sugars into desired end products. For
example, ethanol-producing microorganisms can ferment the sugars to produce bio-
ethanol, while other microorganisms can convert the sugars into various biochemi-
cals, such as organic acids, enzymes, or bio-based materials (Huang et al., 2022a, b).

4 Application of Biomass Waste Management for Treating


Contaminated Wastewater

Water is necessary for all living things (Bhardwaj & Sharma, 2021a, b; Bhardwaj,
2022). Now, the water bodies have been contaminated with the effluent of various
industries, institutes, and residential areas. This effluent comprises decomposed
organic waste that can generate gases with an awful odor in great amounts. This
polluted water contains toxic persistent pollutants that might be hazardous to the
environment (Bhardwaj & Jindal, 2019, 2020, 2022; Bhardwaj et al., 2021). The
usage of enzymes for the treatment of wastewater reduces the toxicity of water and
makes the water fit to reusage.
Most of the pollutants act as substrates for certain enzymes. By the action of
enzymes, the toxic pollutants are fragmented into smaller fragments that no longer
pose a danger to the surroundings. Enzymes can be introduced directly or even
mixed with microbes. Entire body of plants or their tissue culture that contains the
enzymes in their normal form can be deployed in water bodies. BIO-CAT’s enzymes
are effectively used to treat organic waste (Sahal et al., 2023).
Certain enzymes are required to catalyze specific reactions in specified concen-
trations. Oxidoreductases, peroxidases, and oxygenases are used for the removal of
inorganic contaminants such as biphenols, chlorophenols, benzidines, methylated
phenols, phenols, heterocyclic aromatic compounds, and anilines, while lipase, ure-
ase, amylase, xylanases, cellulase, and protease are used for the removal of organic
pollutants. These enzymes are called septic tank enzymes and are utilized for the
treatment of wastewater from sewage. Oxygenases have been used for the treatment
of water bodies that are poisoned by the presence of fossil fuels. Peroxidases are
utilized for the treatment of water that is contaminated with dyes, phenols, and
hydrogen peroxide. Polyphenol oxidases are used in wastewater treatment plants for
the removal of phenol from wastewater and can further be divided into tyrosinase
and laccase.
Joutey et al. (2013) studied the degradation of persistent organic pollutants
(POPs) by using enzymes from microbes and stated that this technique is
environment-­friendly, cost-effective, and innovative. Enzymes, for example, cyto-
chrome P450s laccases, hydrolases, dehalogenases, dehydrogenases, proteases, and
lipases, are involved in the degradation of harmful chemicals (Bhandari et al., 2021).
230 P. Rath et al.

Cytochrome P450 is responsible for synthesizing natural products in living organ-


isms, and it is used in biotransformation of poisonous chemicals present in our
ecosystem (Li et al., 2020). This enzyme has an intrinsic capability to degrade xeno-
biotics (Anzenbacher & Anzenbacherova, 2001; Chakraborty & Das, 2016). P450
BM3 is used in the degradation of several organic pollutants using Pt/TiO2-Cu under
solar radiation and is formed from E. coli BL21 (Awad & Mohamed, 2019).
Laccase has been observed to possess catalytic capability to degrade the aro-
matic amine and phenolic compounds (Chandra & Chowdhary, 2015). It degrades
the PAHs to carbon dioxide (CO2) (Khlifi et al. 2010) and converts acenaphthylene
to 1,8-naphthalic acid and 1,2-acenapthalenedione (Madhavi & Lele, 2009). It can
detoxify dyes produced by the textile industry (Sondhi et al., 2015).
Dehalogenases belong to the oxidoreductase family and are used for the degrada-
tion of halogenated compounds. Reductive, oxygenolytic, and hydrolytic mecha-
nisms are used in the breakdown of halogenated compounds (Wang et al., 2018a, b).
Halohydrin dehydrogenase (HHDH) and haloalkane dehydrogenase (HADH) are
used in the degradation of haloalkane and are expressed in E. coli (Xue et al., 2018).
2,4,6-Trichlorophenol reductive dehalogenase dechlorinates pentachlorophenol
(PCP) into 3-chlorophenol and is isolated from Desulfitobacterium frappieri PCP-1
(Boyer et al., 2003). Alcohol dehydrogenase catalyzes the conversion process of
alcohol into ketone or aldehyde. It also catalyzed NAD(P)+-dependent oxidation of
aldehydes into carboxylic acids.
Polyethylene glycol dehydrogenase (PEGDH) degrades the pollutants that are
released from industries. The enzyme NAD+-dependent polypropylene glycol
dehydrogenase (PPGDH) is isolated from the species Stenotrophomonas malto-
philia and plays a role in oxidization of secondary alcohols (Tachibana et al., 2008).
Polyvinyl alcohol dehydrogenase (PVADH) degrades polyvinyl alcohol, which is a
water-soluble contaminant (Hirota-Mamoto et al., 2006). Aldehyde dehydrogenase
(ADH) degrades aromatic compounds and catalyzes the conversion of 1-hydroxy-­2-
naphthaldehyde to 1-hydroxy-2-naphthoic (Ji et al., 2020).
Hydrolases reduce pollutant toxicity by degrading the larger molecules to smaller
molecules. Hydrolytic enzymes such as amylases, proteases, lipases, esterases,
nitrilases, cellulases, cutinase, and peroxidases can be used in the degradation of
insecticides and oil-contaminated soils. They are used in biomedical sciences and
chemical industries (Kumar & Sharma, 2019). Hydrolases such as parathion hydro-
lase or carbamate, which are isolated from Pseudomonas, Achromobacter,
Flavobacterium, Bacillus cereus, and Nocardia, have been used for converting
diazinon, carbofuran, coumaphos, and carbaryl/parathion through the method of
hydrolysis.
Organophosphate pesticides (OPPs) degrade through hydrolysis of P-O-alkyl
and P-O-aryl bonds (Singh, 2014). Malathion is degraded by the action of
Brevibacillus sp., Alicyclobacillus tengchongensis, Bacillus cereus, and Bacillus
licheniformis (Littlechild, 2015). OP acid anhydrolase enzymes, methyl parathion
hydrolases (MPHs), and OP hydrolases are used for the degradation of organophos-
phates (OPs) (Schenk et al., 2016). Cutinase is used for the degradation of polycap-
rolactone, and it is isolated from the bacterial species Fusarium solani f. pisi (Singh
Significance of Enzymatic Actions in Biomass Waste Management: Challenges… 231

et al., 2016). Protease belongs to the hydrolase family and is isolated from
Amycolatopsis sp., Aspergillus sp., and Bacillus sp. They are low cost, catalyze
enzymes, and are used in industries such as food, leather, and wastewater treatment
(Kumar & Sharma, 2019). They are capable of the degradation of
poly(hydroxybutyrate) (PHB) depolymerase ß-ester bonds, lipase γ-ω bonds, and
α-ester bonds (Haider et al., 2019). They convert marine crustacean wastes and
keratinous wastes into useful products. Keratinase is isolated from Stenotrophomonas
maltophilia KB13 and used in the degradation of chicken feathers (Bhange et al.,
2016). Keratinase is used in the leather industry, and it reduces the chemical (CaO
and Na2S) load in wastewater (Akhter et al., 2020). The enzyme chitinase is isolated
from Bacillus subtilis and degrades crystalline chitin into N-acetyl-D-glucosamine
(Wang et al., 2018a). Pseudomonas fluorescens degrades polyurethane (PU) within
4–5 days using the enzyme protease.
Lipases are a well-known biocatalyst and degrade lipids (Casas-Godoy et al.,
2012). They are used in the degradation of contaminants such as petroleum, resi-
dues of oil, greasy effluents, and oil spills (Basheer et al., 2011; Casas-Godoy et al.,
2012; Hassan et al., 2018). Lipases, which are isolated from the species of
Pseudomonas, are used for the degradation of industrial waste oil (Amara & Salem,
2009). Crude lipase, which is formed from the species Bacillus subtilis, is used in
soap industries to reduce phosphate-based chemicals (Saraswat et al. (2017). Lipase
PL, which is isolated from Alcaligenes sp., catalyzes the conversion of poly(L-­
lactide) (PLA) polymers to the monomers. Lipases that are formed from
Lactobacillus plantarum and Lactobacillus brevis are used in the degradation of
polycaprolactone (PCL) and polyester (Wang et al., 2022).

5 Challenges and Future Prospects

Biomass waste treatment using enzymes from microorganisms is an area of ongoing


research and development and has gained attention. Researchers are continuously
researching to improve enzyme efficiency, process condition optimization, develop-
ment of more cost-effective and sustainable methods, etc. (Madhavan et al., 2021).
Utilizing biomass resources and enzyme diversity and specificity have been areas of
active research. One of the main challenges is scaling up the biomass waste treat-
ment from the laboratory to industrial-scale processes (Kalak, 2023). Advancement
in enzyme engineering techniques, protein engineering, and meta-genomics studies
is required for developing enzymes with enhanced catalytic properties, stability, and
specificity (Kate et al., 2022). Biomass waste management using enzymes could
also be integrated with biorefineries, together aiming to produce multiple valuable
products that are more sustainable and economically viable such as biofuels, bio-­
based chemicals, and materials (Leong et al., 2021). Analyzing the properties of
biomass waste materials would be helpful for understanding the appropriate meth-
ods to utilize them for energy, thereby reducing the consumption of fossil fuels in
the future (Savla et al., 2021).
232 P. Rath et al.

6 Conclusion

There is a need for continuous and vigorous research on the usage of enzymes for
the treatment of waste. Many industries such as textiles, agro, food and beverages,
solid-waste treatment, and sugar mill are using enzymes for the treatment of wastes.
The use of enzymes is a cost-effective and sustainable technique for the treatment
of waste. The usage of enzymes offers many advantages in biomass waste manage-
ment strategies. Enzymatic processes are generally environmentally friendly, highly
specific in their actions, operate under mild conditions, and produce minimal waste
as compared to traditional chemical methods. Furthermore, enzymes are produced
through sustainable and renewable sources, such as microbial cultures. The applica-
tion of enzymes for biomass waste management on a wide-scale/industrial level has
challenges. Overall, the application of enzymes for biomass waste management rep-
resents a promising approach toward achieving a more sustainable and circular
economy.
Acknowledgments The authors thank Amity University for providing the platform to do
this study.
Statement and Declarations:
Conflict of Interest: The authors declare that they have no competing financial interest or personal
relationship that could have appeared to influence the work reported in this chapter.
Author Contribution: All authors have equal contribution.
Data Availabiity: Not applicable.

References

Akcil, A., Karahan, A. G., Ciftci, H., & Sagdic, O. (2003). Biological treatment of cyanide by
natural isolated bacteria (Pseudomonas sp.). Minerals Engineering, 16(7), 643–649.
Akhter, M., Wal Marzan, L., Akter, Y., & Shimizu, K. (2020). Microbial bioremediation of feather
waste for keratinase production: An outstanding solution for leather dehairing in tanneries.
Microbiology Insights, 13, 1178636120913280.
Alam, S., Bhardwaj, L. K., Mallick, R., & Rai, S. (2023). Estimation of heavy metals and fluo-
ride ion in vegetables grown nearby the stretch of River Yamuna, Delhi (NCR), India. Indian
Journal of Environmental Protection, 43(1), 64–73.
Amara, A. A., & Salem, S. R. (2009). Degradation of castor oil and lipase production by
Pseudomonas aeruginosa. American-Eurasian Journal of Agricultural & Environmental
Sciences, 5(4), 556–563.
Anzenbacher, P., & Anzenbacherova, E. (2001). Cytochromes P450 and metabolism of xenobiot-
ics. Cellular and Molecular Life Sciences: CMLS, 58, 737–747.
Arslan, N. P., Yazici, A., Komesli, S., Esim, N., & Ortucu, S. (2021). Direct conversion of waste
loquat kernels to pigments using Monascus purpureus ATCC16365 with proteolytic and amy-
lolytic activity. Biomass Conversion and Biorefinery, 11, 2191–2199.
Awad, G., & Mohamed, E. F. (2019). Immobilization of P450 BM3 monooxygenase on hollow
nanosphere composite: Application for degradation of organic gases pollutants under solar
radiation lamp. Applied Catalysis B: Environmental, 253, 88–95.
Significance of Enzymatic Actions in Biomass Waste Management: Challenges… 233

Bansal, N., & Kanwar, S. S. (2013). Peroxidase (s) in environment protection. The Scientific World
Journal, 2013, 1.
Basheer, S. M., Chellappan, S., Beena, P. S., Sukumaran, R. K., Elyas, K. K., & Chandrasekaran,
M. (2011). Lipase from marine Aspergillus awamori BTMFW032: Production, partial purifica-
tion and application in oil effluent treatment. New Biotechnology, 28(6), 627–638.
Bayramoglu, G., Akbulut, A., & Arica, M. Y. (2013). Immobilization of tyrosinase on modified
diatom biosilica: Enzymatic removal of phenolic compounds from aqueous solution. Journal
of Hazardous Materials, 244, 528–536.
Bhandari, S., Poudel, D. K., Marahatha, R., Dawadi, S., Khadayat, K., Phuyal, S., et al. (2021).
Microbial enzymes used in bioremediation. Journal of Chemistry, 2021, 1–17.
Bhange, K., Chaturvedi, V., & Bhatt, R. (2016). Feather degradation potential of Stenotrophomonas
maltophilia KB13 and feather protein hydrolysate (FPH) mediated reduction of hexavalent
chromium. 3 Biotech, 6, 1–9.
Bhardwaj, L. K. (2022). Evaluation of bis (2-ethylhexyl) phthalate (DEHP) in the PET bottled
mineral water of different brands and impact of heat by GC–MS/MS. Chemistry Africa, 5(4),
929–942.
Bhardwaj, L. K., & Jindal, T. (2019). Contamination of lakes in Broknes peninsula, East Antarctica
through the pesticides and PAHs. Asian Journal of Chemistry, 31(7), 1574–1580.
Bhardwaj, L. K., & Jindal, T. (2020). Persistent organic pollutants in lakes of Grovnes Peninsula at
Larsemann Hill area, East Antarctica. Earth Systems and Environment, 4, 349–358.
Bhardwaj, L. K., & Jindal, T. (2022). Polar ecotoxicology: Sources and toxic effects of pollutants.
In New frontiers in environmental toxicology (pp. 9–14). Springer.
Bhardwaj, L. K., & Sharma, A. (2021a). Microplastics (MPs) in drinking water: Uses, sources
& transport. In Futuristic trends in agriculture engineering & food science. IIP Proceedings.
https://doi.org/10.20944/preprints202104.0498.v1
Bhardwaj, L. K., & Sharma, A. (2021b). Estimation of physico-chemical, trace metals, microbio-
logical and phthalate in PET bottled water. Chemistry Africa, 4(4), 981–991.
Bhardwaj, L. K., Sharma, S., & Jindal, T. (2021). Occurrence of polycyclic aromatic hydrocarbons
(PAHs) in the lake water at Grovnes Peninsula Over East Antarctica. Chemistry Africa, 4,
965–980.
Bhardwaj, L. K., Kumar, D., & Kumar, A. (2023a). Phytoremediation potential of Ocimum sanc-
tum: A sustainable approach for remediation of heavy metals. In Phytoremediation potential
of medicinal and aromatic plants. Taylor & Francis Group. https://doi.org/10.20944/pre-
prints202308.0593.v1
Bhardwaj, L. K., Sharma, S., & Jindal, T. (2023b). Estimation of physico-chemical and heavy met-
als in the lakes of Grovnes & Broknes Peninsula, Larsemann Hill, East Antarctica. Chemistry
Africa, 6, 1–18.
Bilal, M., & Iqbal, H. M. (2021). Ligninolysis potential of ligninolytic enzymes: A green and
sustainable approach to bio-transform lignocellulosic biomass into high-value entities. In
Alternative energy resources: The way to a sustainable modern society (pp. 151–171). Springer.
Boyer, A., Pagé-BéLanger, R., Saucier, M., Villemur, R., Lépine, F., Juteau, P., & Beaudet,
R. (2003). Purification, cloning and sequencing of an enzyme mediating the reductive dechlo-
rination of 2, 4, 6-trichlorophenol from Desulfitobacterium frappieri PCP-1. Biochemical
Journal, 373(1), 297–303.
Casas-Godoy, L., Duquesne, S., Bordes, F., Sandoval, G., & Marty, A. (2012). Lipases: An
overview. In Lipases and phospholipases: Methods and protocols (pp. 3–30). https://doi.
org/10.1007/978-­1-­4939-­8672-­9_1
Castro, A. M., Castilho, L. R., & Freire, D. M. (2011). An overview on advances of amylases
production and their use in the production of bioethanol by conventional and non-conventional
processes. Biomass Conversion and Biorefinery, 1, 245–255.
Chakraborty, J., & Das, S. (2016). Molecular perspectives and recent advances in microbial reme-
diation of persistent organic pollutants. Environmental Science and Pollution Research, 23,
16883–16903.
234 P. Rath et al.

Champagne, P., & Li, C. (2009). Enzymatic hydrolysis of cellulosic municipal wastewater treat-
ment process residuals as feedstocks for the recovery of simple sugars. Bioresource Technology,
100(23), 5700–5706.
Chandra, R., & Chowdhary, P. (2015). Properties of bacterial laccases and their application in
bioremediation of industrial wastes. Environmental Science: Processes & Impacts, 17(2),
326–342.
Chaudhuri, G., Dey, P., Dalal, D., Venu-Babu, P., & Thilagaraj, W. R. (2013). A novel approach to
precipitation of heavy metals from industrial effluents and single-ion solutions using bacterial
alkaline phosphatase. Water, Air, & Soil Pollution, 224, 1–11.
Chu, R., Li, S., Zhu, L., Yin, Z., Hu, D., Liu, C., & Mo, F. (2021). A review on co-cultivation of
microalgae with filamentous fungi: Efficient harvesting, wastewater treatment and biofuel pro-
duction. Renewable and Sustainable Energy Reviews, 139, 110689.
Chung, D., Cha, M., Guss, A. M., & Westpheling, J. (2014). Direct conversion of plant biomass
to ethanol by engineered Caldicellulosiruptor bescii. Proceedings of the National Academy of
Sciences, 111(24), 8931–8936.
Coughlin, R. W., & Charles, M. (2019). Applications of lactase and immobilized lactase. In
Immobilized enzymes for food processing (pp. 153–173). CRC Press.
Dai, Z., Liu, L., Duan, H., Li, B., Tang, X., Wu, X., et al. (2023). Improving sludge dewaterability
by free nitrous acid and lysozyme pretreatment: Performances and mechanisms. Science of the
Total Environment, 855, 158648.
Dashtban, M., Maki, M., Leung, K. T., Mao, C., & Qin, W. (2010). Cellulase activities in biomass
conversion: Measurement methods and comparison. Critical Reviews in Biotechnology, 30(4),
302–309.
Demirbas, M. F. (2009). Biorefineries for biofuel upgrading: A critical review. Applied Energy,
86, S151–S161.
Ebbs, S. (2004). Biological degradation of cyanide compounds. Current Opinion in Biotechnology,
15(3), 231–236.
Falade, A. O., Nwodo, U. U., Iweriebor, B. C., Green, E., Mabinya, L. V., & Okoh, A. I. (2017).
Lignin peroxidase functionalities and prospective applications. Microbiology Open,
6(1), e00394.
Feng, Y. S., Chen, P. C., Wen, F. S., Hsiao, W. Y., & Lee, C. M. (2008). Nitrile hydratase from
Mesorhizobium sp. F28 and its potential for nitrile biotransformation. Process Biochemistry,
43(12), 1391–1397.
Ferreira, J. A., Mahboubi, A., Lennartsson, P. R., & Taherzadeh, M. J. (2016). Waste biorefin-
eries using filamentous ascomycetes fungi: Present status and future prospects. Bioresource
Technology, 215, 334–345.
Guo, H., Zhao, Y., Chang, J. S., & Lee, D. J. (2023). Enzymes and enzymatic mechanisms in
enzymatic degradation of lignocellulosic biomass: A mini-review. Bioresource Technology,
367, 128252.
Haider, T. P., Völker, C., Kramm, J., Landfester, K., & Wurm, F. R. (2019). Plastics of the future?
The impact of biodegradable polymers on the environment and on society. Angewandte Chemie
International Edition, 58(1), 50–62.
Hassan, S. W., Abd El Latif, H. H., & Ali, S. M. (2018). Production of cold-active lipase by free
and immobilized marine Bacillus cereus HSS: Application in wastewater treatment. Frontiers
in Microbiology, 9, 2377.
Hirota-Mamoto, R., Nagai, R., Tachibana, S., Yasuda, M., Tani, A., Kimbara, K., & Kawai,
F. (2006). Cloning and expression of the gene for periplasmic poly (vinyl alcohol) dehydroge-
nase from Sphingomonas sp. strain 113P3, a novel-type quinohaemoprotein alcohol dehydro-
genase. Microbiology, 152(7), 1941–1949.
Huang, C., Li, R., Tang, W., Zheng, Y., & Meng, X. (2022a). Improve enzymatic hydrolysis of
lignocellulosic biomass by modifying lignin structure via sulfite pretreatment and using lignin
blockers. Fermentation, 8(10), 558.
Significance of Enzymatic Actions in Biomass Waste Management: Challenges… 235

Huang, C., Jiang, X., Shen, X., Hu, J., Tang, W., Wu, X., et al. (2022b). Lignin-enzyme interaction:
A roadblock for efficient enzymatic hydrolysis of lignocellulosics. Renewable and Sustainable
Energy Reviews, 154, 111822.
Ji, D., Mao, Z., He, J., Peng, S., & Wen, H. (2020). Characterization and genomic function analy-
sis of phenanthrene-degrading bacterium Pseudomonas sp. Lphe-2. Journal of Environmental
Science and Health, Part A, 55(5), 549–562.
Joutey, N. T., Bahafid, W., Sayel, H., & El Ghachtouli, N. (2013). Biodegradation: Involved micro-
organisms and genetically engineered microorganisms. Biodegradation-Life of Science, 1,
289–320.
Kalak, T. (2023). Potential use of industrial biomass waste as a sustainable energy source in the
future. Energies, 16(4), 1783.
Karn, S. K., & Kumar, A. (2015). Hydrolytic enzyme protease in sludge: recovery and its applica-
tion. Biotechnology and Bioprocess Engineering, 20, 652–661.
Kate, A., Sahu, L. K., Pandey, J., Mishra, M., & Sharma, P. K. (2022). Green catalysis for chemi-
cal transformation: The need for the sustainable development. Current Research in Green and
Sustainable Chemistry, 5, 100248.
Khan, M. N., Luna, I. Z., Islam, M. M., Sharmeen, S., Salem, K. S., Rashid, T. U., et al. (2016).
Cellulase in waste management applications. In New and future developments in microbial
biotechnology and bioengineering (pp. 237–256). Elsevier.
Khatami, S. H., Vakili, O., Movahedpour, A., Ghesmati, Z., Ghasemi, H., & Taheri-Anganeh,
M. (2022). Laccase: Various types and applications. Biotechnology and Applied Biochemistry,
69(6), 2658–2672.
Khlifi, R., Belbahri, L., Woodward, S., Ellouz, M., Dhouib, A., Sayadi, S., & Mechichi, T. (2010).
Decolourization and detoxification of textile industry wastewater by the laccase-mediator sys-
tem. Journal of Hazardous Materials, 175(1–3), 802–808.
Knecht, L. E., Veljkovic, M., & Fieseler, L. (2020). Diversity and function of phage encoded
depolymerases. Frontiers in Microbiology, 10, 2949.
Kohli, P., Kalia, M., & Gupta, R. (2015). Pectin methylesterases: A review. Journal of Bioprocessing
& Biotechniques, 5(5), 1.
Kolusheva, T., & Marinova, A. (2007). A study of the optimal conditions for starch hydroly-
sis through thermostable α-amylase. Journal of the University of Chemical Technology and
Metallurgy, 42(1), 93–96.
Kumar, A., & Sharma, S. (2019). Microbes and enzymes in soil health and bioremediation
(pp. 353–366). Springer.
Leong, H. Y., Chang, C. K., Khoo, K. S., Chew, K. W., Chia, S. R., Lim, J. W., et al. (2021).
Waste biorefinery towards a sustainable circular bioeconomy: A solution to global issues.
Biotechnology for Biofuels, 14(1), 1–15.
Li, Z., Jiang, Y., Guengerich, F. P., Ma, L., Li, S., & Zhang, W. (2020). Engineering cytochrome
P450 enzyme systems for biomedical and biotechnological applications. Journal of Biological
Chemistry, 295(3), 833–849.
Li, S., Zhong, L., Wang, H., Li, J., Cheng, H., & Ma, Q. (2021). Process optimization of poly-
phenol oxidase immobilization: Isotherm, kinetic, thermodynamic and removal of phenolic
compounds. International Journal of Biological Macromolecules, 185, 792–803.
Littlechild, J. A. (2015). Archaeal enzymes and applications in industrial biocatalysts. Archaea,
2015, 1.
Liu, L., Bilal, M., Duan, X., & Iqbal, H. M. (2019). Mitigation of environmental pollution by
genetically engineered bacteria – Current challenges and future perspectives. Science of the
Total Environment, 667, 444–454.
Madhavan, A., Arun, K. B., Binod, P., Sirohi, R., Tarafdar, A., Reshmy, R., et al. (2021). Design of
novel enzyme biocatalysts for industrial bioprocess: Harnessing the power of protein engineer-
ing, high throughput screening and synthetic biology. Bioresource Technology, 325, 124617.
Madhavi, V., & Lele, S. S. (2009). Laccase: Properties and applications. BioResources, 4(4), 1694.
236 P. Rath et al.

Mahapatra, S., Kumar, D., Singh, B., & Sachan, P. K. (2021). Biofuels and their sources of produc-
tion: A review on cleaner sustainable alternative against conventional fuel, in the framework of
the food and energy nexus. Energy Nexus, 4, 100036.
Mander, P., Cho, S. S., Choi, Y. H., Panthi, S., Choi, Y. S., Kim, H. M., & Yoo, J. C. (2016).
Purification and characterization of chitinase showing antifungal and biodegradation properties
obtained from Streptomyces anulatus CS242. Archives of Pharmacal Research, 39, 878–886.
Méndez-Líter, J. A., de Eugenio, L. I., Nieto-Domínguez, M., Prieto, A., & Martínez, M. J. (2021).
Hemicellulases from Penicillium and Talaromyces for lignocellulosic biomass valorization: A
review. Bioresource Technology, 324, 124623.
Muthukumarasamy, N. P., Jackson, B., Joseph Raj, A., & Sevanan, M. (2015). Production of extra-
cellular laccase from Bacillus subtilis MTCC 2414 using agroresidues as a potential substrate.
Biochemistry Research International, 2015, 765190.
Nicell, J. A. (2003). Enzymatic treatment of waters and wastes. In Chemical degradation methods
for wastes and pollutants (pp. 395–441). CRC Press.
Nimkande, V. D., & Bafana, A. (2022). A review on the utility of microbial lipases in wastewater
treatment. Journal of Water Process Engineering, 46, 102591.
Pandey, K., Singh, B., Pandey, A. K., Badruddin, I. J., Pandey, S., Mishra, V. K., & Jain, P. A. (2017).
Application of microbial enzymes in industrial waste water treatment. International Journal of
Current Microbiology and Applied Sciences, 6(8), 1243–1254.
Pérez-Prior, M. T., González-Sánchez, M. I., & Valero, E. (2014). Removal of aromatic compounds
from wastewater by hemoglobin soluble and immobilized on Eupergit® CM. Environmental
Engineering & Management Journal (EEMJ), 13(10), 2459.
Phale, P. S., Sharma, A., & Gautam, K. (2019). Microbial degradation of xenobiotics like aromatic
pollutants from the terrestrial environments. In Pharmaceuticals and personal care products:
Waste management and treatment technology (pp. 259–278). Butterworth-Heinemann.
Presečki, A. V., Blažević, Z. F., & Vasić-Rački, Đ. (2013). Complete starch hydrolysis by the
synergistic action of amylase and glucoamylase: Impact of calcium ions. Bioprocess and
Biosystems Engineering, 36, 1555–1562.
Sahal, S., Khaturia, S., & Joshi, N. (2023). Application of microbial enzymes in wastewater treat-
ment. In Genomics approach to bioremediation: Principles, tools, and emerging technologies
(pp. 209–227). Wiley.
Sandhya, S. (2010). Biodegradation of azo dyes under anaerobic condition: Role of azoreductase.
In Biodegradation of azo dyes (pp. 39–57). https://doi.org/10.1007/698_2009_43
Saraswat, R., Verma, V., Sistla, S., & Bhushan, I. (2017). Evaluation of alkali and thermotolerant
lipase from an indigenous isolated Bacillus strain for detergent formulation. Electronic Journal
of Biotechnology, 30, 33–38.
Savla, N., Pandit, S., Mathuriya, A. S., Gupta, P. K., Khanna, N., Babu, R. P., & Kumar, S. (2021).
Recent advances in bioelectricity generation through the simultaneous valorization of lig-
nocellulosic biomass and wastewater treatment in microbial fuel cell. Sustainable Energy
Technologies and Assessments, 48, 101572.
Schenk, G., Mateen, I., Ng, T. K., Pedroso, M. M., Mitić, N., Jafelicci, M., Jr., et al. (2016).
Organophosphate-degrading metallohydrolases: Structure and function of potent catalysts for
applications in bioremediation. Coordination Chemistry Reviews, 317, 122–131.
Sharma, S. (2012). Bioremediation: Features, strategies and applications. Asian Journal of
Pharmacy and Life Science, 2231, 4423.
Shekher, R., Sehgal, S., Kamthania, M., & Kumar, A. (2011). Laccase: Microbial sources, produc-
tion, purification, and potential biotechnological applications. Enzyme Research, 2011, 217861.
Singh, B. (2014). Review on microbial carboxylesterase: General properties and role in organo-
phosphate pesticides degradation. Biochemistry and Molecular Biology, 2, 1–6.
Singh, R., Kumar, M., Mittal, A., & Mehta, P. K. (2016). Microbial enzymes: Industrial progress
in 21st century. 3 Biotech, 6, 1–15.
Siqueira, J. G. W., Rodrigues, C., de Souza Vandenberghe, L. P., Woiciechowski, A. L., & Soccol,
C. R. (2020). Current advances in on-site cellulase production and application on lignocellu-
losic biomass conversion to biofuels: A review. Biomass and Bioenergy, 132, 105419.
Significance of Enzymatic Actions in Biomass Waste Management: Challenges… 237

Sjoblad, R. D., & Bollag, J. M. (2021). Oxidative coupling of aromatic compounds by enzymes
from soil microorganisms. In Soil biochemistry (pp. 113–152). CRC Press.
Soares, P. R. S., Birolli, W. G., Ferreira, I. M., & Porto, A. L. M. (2021). Biodegradation path-
way of the organophosphate pesticides chlorpyrifos, methyl parathion and profenofos by the
marine-derived fungus Aspergillus sydowii CBMAI 935 and its potential for methylation reac-
tions of phenolic compounds. Marine Pollution Bulletin, 166, 112185.
Sondhi, S., Sharma, P., George, N., Chauhan, P. S., Puri, N., & Gupta, N. (2015). An extracellular
thermo-alkali-stable laccase from Bacillus tequilensis SN4, with a potential to biobleach soft-
wood pulp. 3 Biotech, 5, 175–185.
Tachibana, S., Naka, N., Kawai, F., & Yasuda, M. (2008). Purification and characterization of
cytoplasmic NAD+-dependent polypropylene glycol dehydrogenase from Stenotrophomonas
maltophilia. FEMS Microbiology Letters, 288(2), 266–272.
Toesch, M., Schober, M., & Faber, K. (2014). Microbial alkyl-and aryl-sulfatases: Mechanism,
occurrence, screening and stereoselectivities. Applied Microbiology and Biotechnology, 98,
1485–1496.
Wang, D., Li, A., Han, H., Liu, T., & Yang, Q. (2018a). A potent chitinase from Bacillus subtilis
for the efficient bioconversion of chitin-containing wastes. International Journal of Biological
Macromolecules, 116, 863–868.
Wang, Y., Feng, Y., Cao, X., Liu, Y., & Xue, S. (2018b). Insights into the molecular mechanism
of dehalogenation catalyzed by D-2-haloacid dehalogenase from crystal structures. Scientific
Reports, 8(1), 1–9.
Wang, Y., Huang, J., Liang, X., Wei, M., Liang, F., Feng, D., et al. (2022). Production and waste
treatment of polyesters: Application of bioresources and biotechniques. Critical Reviews in
Biotechnology, 43, 1–18.
Xue, F., Ya, X., Tong, Q., Xiu, Y., & Huang, H. (2018). Heterologous overexpression of
Pseudomonas umsongensis halohydrin dehalogenase in Escherichia coli and its application in
epoxide asymmetric ring opening reactions. Process Biochemistry, 75, 139–145.
Yadav, S., Yadav, P. K., Yadav, D., & Yadav, K. D. S. (2009). Pectin lyase: A review. Process
Biochemistry, 44(1), 1–10.
Zanuso, E., Gomes, D. G., Ruiz, H. A., Teixeira, J. A., & Domingues, L. (2021). Enzyme immobi-
lization as a strategy towards efficient and sustainable lignocellulosic biomass conversion into
chemicals and biofuels: Current status and perspectives. Sustainable Energy & Fuels, 5(17),
4233–4247.
Zhang, H., Han, L., & Dong, H. (2021). An insight to pretreatment, enzyme adsorption and enzy-
matic hydrolysis of lignocellulosic biomass: Experimental and modeling studies. Renewable
and Sustainable Energy Reviews, 140, 110758.
Zhao, S., Xu, W., Zhang, W., Wu, H., Guang, C., & Mu, W. (2021). In-depth biochemical identifica-
tion of a novel methyl parathion hydrolase from Azohydromonas australica and its high effec-
tiveness in the degradation of various organophosphorus pesticides. Bioresource Technology,
323, 124641.
Bioeconomy: A Sustainable Approach
for Biomass Waste Management

Rwitabrata Mallick, Kuldip Dwivedi, and Swapnil Rai

1 Introduction

The bioeconomy refers to an economic system that utilizes renewable biological


resources to produce various goods and services. It encompasses the sustainable use
of crops, forests, marine resources, and microorganisms to generate food, energy,
materials, and other bio-based products. The bioeconomy aims to replace or reduce
reliance on fossil fuels and non-renewable resources by transitioning to a more sus-
tainable model based on biological inputs. It emphasizes the efficient utilization of
biomass and waste streams, promoting circularity and minimizing environmental
impact. Key sectors within the bioeconomy include agriculture, forestry, fisheries,
bioenergy, bio-based chemicals, biomaterials, and biotechnology. These sectors
employ innovative technologies and processes to convert biological resources into
value-added products and services. The bioeconomy also focuses on fostering sus-
tainable practices, conservation, and biodiversity protection. It aims to balance eco-
nomic growth with social and environmental considerations, striving for a more
sustainable and resilient society.
The development of the bioeconomy offers opportunities for economic growth,
job creation, and regional development, while also addressing challenges such as
climate change, resource scarcity, and waste management. By utilizing renewable
biological resources, the bioeconomy promotes a more sustainable and circular
approach to production and consumption. One of the primary objectives of the bio-
economy is to reduce dependence on fossil-based resources and shift toward renew-
able resources. By doing so, it aims to mitigate the negative environmental impacts

R. Mallick (*) · K. Dwivedi · S. Rai


Department of Environmental Science, Amity School of Life Science, Amity University,
Gwalior, Madhya Pradesh, India
e-mail: rmallick@gwa.amity.edu

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 239
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_13
240 R. Mallick et al.

associated with the extraction and consumption of non-renewable resources, such as


fossil fuels (Bugge et al., 2016).
The bioeconomy promotes the use of sustainable practices and processes that
prioritize environmental stewardship. This includes the responsible management of
renewable biological resources, such as crops, forests, and marine ecosystems, to
ensure their long-term viability and health. It also encourages the adoption of effi-
cient and environmentally friendly technologies and production methods. By
embracing renewable resources and sustainable practices, the bioeconomy aims to
minimize greenhouse gas emissions, reduce waste generation, and conserve natural
resources. This contributes to the preservation of the environment and the mitiga-
tion of climate change impacts.
Additionally, the bioeconomy seeks to support economic growth and develop-
ment. By creating markets for bio-based products and services, it can generate new
business opportunities and job creation. This economic growth is pursued in a man-
ner that is compatible with sustainability principles, considering the social and envi-
ronmental aspects alongside economic prosperity. Ultimately, the bioeconomy
strives to achieve a balance between economic growth, environmental preservation,
and social well-being. By transitioning to renewable resources and embracing sus-
tainable practices, it aims to create a more resilient and prosperous society that can
meet its needs without compromising the well-being of future generations. The bio-
economy encompasses a diverse range of industries that contribute to the sustain-
able utilization of renewable biological resources (Carus & Dammer, 2018).
The bioeconomy involves the use of agricultural resources to produce food, feed,
and various bio-based products. It includes sustainable farming practices, precision
agriculture, and the development of genetically modified crops for enhanced pro-
ductivity and resilience. Agriculture plays a significant role in the bioeconomy. It is
a key sector within the bioeconomy framework as it involves the sustainable pro-
duction of crops, livestock, and other agricultural products for food, feed, and vari-
ous bio-based applications. The bioeconomy promotes sustainable farming practices
that focus on optimizing resource efficiency, reducing environmental impact, and
ensuring the long-term productivity of agricultural systems. This includes practices
such as precision agriculture, organic farming, agroecology, and the use of inte-
grated pest management techniques. Sustainable farming practices are a crucial
component of the bioeconomy and play a significant role in promoting environmen-
tal stewardship, resource efficiency, and long-term agricultural productivity. Organic
farming practices prioritize the use of natural fertilizers, crop rotation, biological
pest control, and the avoidance of synthetic pesticides and genetically modified
organisms (GMOs). Organic farming promotes soil health and biodiversity and
reduces chemical inputs, thereby minimizing environmental impact. Precision agri-
culture utilizes technologies such as GPS, remote sensing, and data analytics to
optimize the use of inputs such as water, fertilizers, and pesticides. By applying
these inputs precisely where and when they are needed, precision agriculture mini-
mizes waste, enhances resource efficiency, and reduces environmental impact
(Dietz et al., 2018).
Bioeconomy: A Sustainable Approach for Biomass Waste Management 241

Conservation agriculture involves practices that minimize soil disturbance, such


as reduced tillage or no-till farming, to protect soil structure and reduce erosion. It
promotes the use of cover crops, crop rotation, and organic mulches to improve soil
health, moisture retention, and nutrient cycling. Agroforestry combines agricultural
crops with trees and shrubs to create integrated and sustainable land-use systems.
Agroforestry provides multiple benefits, such as improving soil fertility, enhancing
biodiversity, sequestering carbon, and providing shade and windbreaks for crops
and livestock. Sustainable farming practices include efficient water management
techniques such as drip irrigation, precision sprinklers, and water recycling. These
methods help conserve water resources, reduce water pollution from agricultural
runoff, and enhance water-use efficiency. Integrated Pest Management (IPM): IPM
is an approach that focuses on using a combination of biological, cultural, and
chemical control methods to manage pests and diseases in a sustainable manner. It
aims to minimize reliance on synthetic pesticides and emphasizes the use of natural
predators, crop rotation, and resistant crop varieties.
Sustainable farming involves optimizing nutrient management through practices
such as soil testing, targeted fertilization, and precision application of fertilizers.
This helps minimize nutrient runoff and water pollution while ensuring that crops
receive the necessary nutrients for healthy growth. Sustainable farming practices
prioritize the preservation and enhancement of on-farm biodiversity. This can be
achieved through the creation of habitat for beneficial insects and wildlife, the
establishment of buffer zones, and the protection of natural areas within and around
agricultural landscapes. By adopting sustainable farming practices, farmers can
enhance the resilience and productivity of their operations while minimizing nega-
tive environmental impacts. These practices contribute to the principles of the bio-
economy by ensuring the sustainable use of agricultural resources, protecting
ecosystems, and promoting the production of environmentally friendly and healthy
food (El-Chichakli et al., 2016).
Agriculture contributes to the bioeconomy by providing bio-based inputs that are
used in the production of various bio-based products. This includes crops grown for
biofuels, biomaterials, and bioplastics, and feedstocks for the biorefinery industry.
Bio-based inputs refer to raw materials or ingredients derived from biological
sources, such as plants, animals, and microorganisms. These inputs are utilized in
various industries and applications, including agriculture, manufacturing, energy
production, and consumer goods. The use of bio-based inputs is driven by the desire
to reduce reliance on fossil fuels, decrease environmental impact, and promote sus-
tainable practices. In agriculture, bio-based inputs can include organic fertilizers,
biopesticides, and biostimulants derived from natural sources. These inputs provide
an alternative to synthetic chemicals, promoting soil health, reducing pollution, and
supporting sustainable farming practices. In manufacturing, bio-based inputs can be
used as alternatives to petroleum-based chemicals and materials. For example, bio-­
based plastics, such as those derived from corn starch or sugarcane, can be used in
packaging materials, consumer goods, and even automotive parts. Bio-based sol-
vents, adhesives, and lubricants are also gaining popularity as sustainable
alternatives.
242 R. Mallick et al.

The energy sector is another area where bio-based inputs play a significant role.
Biomass, such as agricultural waste, forest residues, and dedicated energy crops,
can be converted into biofuels such as bioethanol and biodiesel. These renewable
fuels offer a lower carbon footprint compared to fossil fuels and can be used in
transportation and power generation. Additionally, bio-based inputs are used in the
production of various consumer goods. Natural fibers such as cotton, hemp, and
bamboo can be utilized in textiles and clothing manufacturing, providing an eco-­
friendly alternative to synthetic fibers. Personal care products, cosmetics, and clean-
ing agents also often incorporate bio-based ingredients derived from plants and
natural extracts (McCormick & Kautto, 2013).
The development and utilization of bio-based inputs contribute to the transition
toward a more sustainable and circular economy. By relying on renewable resources
and reducing dependence on fossil fuels, these inputs offer the potential for
decreased environmental impact and a more sustainable future. Certain crops, such
as corn, sugarcane, and switchgrass, are cultivated specifically for bioenergy pro-
duction. These crops are used to produce biofuels such as ethanol and biodiesel,
contributing to the renewable energy sector of the bioeconomy. Bioenergy crops,
also known as energy crops or dedicated energy crops, are specific plants cultivated
for the purpose of producing biomass that can be converted into renewable energy.
These crops are grown specifically to be used as a feedstock for biofuel production
or for direct combustion in bioenergy systems.
Corn is one of the most used bioenergy crops. It is primarily used to produce
bioethanol, which can be blended with gasoline as a renewable fuel source. Corn
stover (the leaves, stalks, and cobs remaining after grain harvest) can also be used
for cellulosic ethanol production or as biomass for other bioenergy applications
(Birner, 2018). Sugarcane is a highly efficient bioenergy crop used mainly to pro-
duce bioethanol. Brazil is a major producer of sugarcane ethanol, utilizing the
sucrose-rich juice extracted from the cane stalks. Sugarcane residues, known as
bagasse, can also be burned to generate heat and electricity.
Switchgrass is a perennial grass native to North America and is often used as a
dedicated energy crop. It has high biomass yield potential and can grow on marginal
lands. Switchgrass is suitable for cellulosic ethanol production, where the cellulose
content of the plant is converted into ethanol through various biochemical or ther-
mochemical processes. Miscanthus is another perennial grass that has gained atten-
tion as an energy crop. It has high biomass productivity and low nutrient requirements
and can grow in a wide range of climates. Miscanthus can be used as a feedstock for
bioenergy production, including cellulosic ethanol or as a solid biomass fuel for
heating or electricity generation. Willow and poplar trees are fast-growing woody
crops that can be harvested for bioenergy purposes. They can be coppiced, where
the stems are cut back to ground level and regrow, allowing for multiple harvests
over several years. Willow and poplar can be used to produce biomass pellets, bio-
char, or as a feedstock for biogas production. Jatropha is a tropical oilseed crop that
has gained attention for its potential to produce biodiesel. The seeds of the Jatropha
plant contain oil that can be extracted and converted into a renewable fuel source.
Jatropha can grow on marginal lands and is drought-resistant, making it suitable for
Bioeconomy: A Sustainable Approach for Biomass Waste Management 243

cultivation in arid regions. The choice of bioenergy crop depends on factors such as
climate, soil conditions, regional availability, and the specific bioenergy conversion
technology being used.
Agricultural residues and by-products, such as crop residues, straw, and animal
waste, can be utilized as feedstocks for bioenergy production, biogas generation, or
the production of bio-based chemicals and materials. This helps to reduce waste and
enhance the overall efficiency of agricultural systems. Agricultural residues and by-­
products refer to the leftover materials generated during agricultural processes.
These materials are typically not the primary focus of agricultural production but
can still have various uses and values. Crop residues are the remaining parts of crops
such as stalks, leaves, husks, and stems after the primary harvest. Crop residues can
be used for several purposes, including animal feed, composting, bioenergy produc-
tion (e.g., biofuels and biomass pellets), and as a raw material in various industries.
Livestock production generates significant amounts of manure, which can be uti-
lized as a nutrient-rich fertilizer for crop production. Manure can also be processed
through anaerobic digestion to produce biogas, which is a renewable energy source
(Pfau et al., 2014).
During the processing of agricultural commodities, various by-products are gen-
erated. For example, in the production of fruits and vegetables, trimmings, peels,
and seeds may be considered waste. These by-products can be used for animal feed,
composting, or as ingredients in food processing (e.g., fruit peel extracts). Rice
production generates large quantities of husks, which can be utilized in several
ways. They can serve as a source of energy through combustion or gasification, used
as a raw material for producing bio-based products (e.g., biofuels and building
materials), or as a bedding material for livestock. Bagasse is the fibrous residue left
after sugarcane stalks are crushed to extract juice. It is commonly used as a biofuel
for heat and electricity generation in sugarcane-producing regions. Bagasse can also
be transformed into other value-added products such as paper, board, and biode-
gradable plastics.
Olive oil production generates olive pomace, which consists of crushed olive
skins, pulp, and pits. This by-product can be used for composting, as a source of
bioactive compounds, or processed to obtain olive pomace oil, which is lower in
quality than extra virgin olive oil but still suitable for certain applications. The utili-
zation of agricultural residues and by-products can help reduce waste, create addi-
tional revenue streams, promote sustainability, and contribute to the circular
economy by turning waste into valuable resources. The specific applications and
uses of these residues may vary depending on factors such as local regulations,
available technologies, and market demand.
Biotechnology plays a crucial role in agriculture, enabling the development of
genetically modified crops with improved traits, such as higher yields, resistance to
pests and diseases, and enhanced nutritional content. Biotechnology also supports
the development of advanced breeding techniques and biotech-derived agricultural
inputs, contributing to the bioeconomy’s advancement. Bioeconomy promotes the
concept of circular agriculture, which involves the efficient use of resources and the
integration of different agricultural systems. This includes practices such as nutrient
244 R. Mallick et al.

recycling, composting, and using organic waste as fertilizer, contributing to the cir-
cular economy principles. Circular agriculture, also known as regenerative agricul-
ture or sustainable agriculture, is an approach to farming that seeks to create a
closed-loop system where resources are conserved, waste is minimized, and eco-
logical systems are enhanced. It aims to mimic the natural cycles and processes
found in ecosystems while ensuring the production of food, fiber, and other agricul-
tural products (Dietz et al., 2018).
Circular agriculture focuses on minimizing resource inputs and maximizing
resource efficiency. This involves reducing the use of synthetic fertilizers, pesti-
cides, and water and promoting practices that enhance soil health and fertility.
Instead of relying heavily on external inputs, circular agriculture emphasizes the
recycling and reuse of nutrients within the system. It involves practices such as
composting, cover cropping, crop rotation, and the use of animal manure to replen-
ish soil nutrients.
Circular agriculture aims to minimize waste and utilize by-products and residues
effectively. Agricultural residues, food waste, and other organic materials are often
composted or processed through anaerobic digestion to generate renewable energy
or nutrient-rich compost. Circular agriculture recognizes the importance of biodi-
versity in maintaining ecological balance. It emphasizes practices that promote bio-
diversity, such as maintaining hedgerows, creating wildlife habitats, and integrating
diverse crop rotations and intercropping systems. Circular agriculture aims to miti-
gate climate change and adapt to its impacts. It promotes practices that sequester
carbon in the soil, such as agroforestry, conservation tillage, and the use of cover
crops. It also focuses on water management and efficient irrigation techniques to
address water scarcity issues.
Circular agriculture emphasizes local production and distribution systems,
reducing the reliance on long-distance transportation and supporting local econo-
mies. This helps to reduce carbon emissions associated with the transportation of
agricultural products. Circular agriculture is seen as a more sustainable alternative
to conventional farming practices. It can contribute to soil health and fertility, reduce
environmental pollution, conserve resources, and enhance the resilience of farming
systems in the face of climate change. By adopting circular agriculture principles,
farmers can improve the long-term viability and sustainability of their operations
while promoting ecosystem health and biodiversity. With the implementation of
incorporating sustainable practices, leveraging biotechnology, and exploring the
potential of agricultural resources, the bioeconomy seeks to enhance the productiv-
ity, resilience, and environmental sustainability of the agricultural sector, while also
generating economic value through the production of bio-based products (Patermann
& Aguilar, 2018).
The bioeconomy promotes the sustainable management of forests to produce
timber, pulp, and paper, as well as bio-based materials such as wood-based compos-
ites. It also explores the potential of forest biomass for bioenergy production. The
bioeconomy recognizes the importance of sustainable fisheries and responsible
aquaculture practices. It focuses on utilizing fish and marine resources for food
production, as well as extracting valuable compounds for pharmaceuticals,
Bioeconomy: A Sustainable Approach for Biomass Waste Management 245

nutraceuticals, and cosmetics. It encourages the production of renewable energy


from biomass sources. This includes biofuels (such as ethanol and biodiesel), bio-
gas, and solid biomass used for heating, electricity generation, and transportation.
Biotechnology plays a crucial role in the bioeconomy by harnessing the potential
of living organisms, such as microorganisms, plants, and animals, for various appli-
cations. This includes the development of enzymes, bio-based chemicals, bioplas-
tics, and biopharmaceuticals. Biotechnology is a fundamental component of the
bioeconomy, driving innovation and enabling the sustainable use of biological
resources to develop a wide range of products and applications. Here is an overview
of the relationship between biotechnology and the bioeconomy: Biotechnology
involves the manipulation and application of living organisms, such as microorgan-
isms, plants, and animals, to develop new products, processes, and technologies.
This includes genetic engineering, fermentation, and other techniques to modify
organisms for specific purposes. Biotechnology plays a significant role in crop
improvement and agricultural productivity. It enables the development of geneti-
cally modified crops with enhanced traits such as increased yields, improved resis-
tance to pests and diseases, and tolerance to environmental stressors. Biotechnology
also supports the development of advanced breeding techniques, molecular diag-
nostics, and precision agriculture practices.
Industrial biotechnology uses enzymes, microorganisms, and other biologically
derived components to produce bio-based chemicals, materials, and energy. It
enables the production of biofuels, bio-based plastics, enzymes, biopolymers, and
other bio-based products that serve as alternatives to fossil fuel-based counterparts.
Biotechnology is instrumental in the development of biopharmaceuticals, which are
drugs produced using biological systems. This includes the production of therapeu-
tic proteins, antibodies, vaccines, and other bio-based medicines. Biotechnology
also supports advancements in diagnostics, gene therapies, and personalized medi-
cine. It contributes to environmental sustainability by offering solutions for pollu-
tion control, waste management, and resource recovery. For example, bioremediation
uses microorganisms to clean up pollutants in soil and water, while microbial fer-
mentation can convert organic waste into valuable products such as biogas or bio-
fertilizers (Stegmann et al., 2020).
Biotechnology drives continuous research and development, leading to new dis-
coveries, techniques, and applications. This ongoing innovation fuels the expansion
of the bioeconomy by exploring untapped potential in biological resources and find-
ing novel solutions to societal challenges. Biotechnology facilitates collaboration
across sectors within the bioeconomy, bringing together researchers, industry, and
policymakers. It enables the integration of knowledge and expertise from diverse
fields, such as agriculture, healthcare, energy, and materials science, to drive sus-
tainable economic growth. By leveraging biotechnology, the bioeconomy seeks to
harness the capabilities of living organisms and their molecular components to
develop sustainable solutions for various industries. It enables the transition from
fossil-based resources to renewable biological resources, fosters innovation, and
promotes the development of environmentally friendly products and processes
(Zilberman et al., 2013).
246 R. Mallick et al.

The bioeconomy explores the production of sustainable materials from renew-


able resources. This includes bio-based plastics, textiles, packaging materials, and
construction materials that can replace traditional fossil-based counterparts.
Biomaterials are materials derived from renewable biological resources that can be
used in various applications, ranging from medical and healthcare to packaging and
construction. They are a key component of the bioeconomy and play a vital role in
reducing reliance on non-renewable resources and promoting sustainability.
Biomaterials are materials that are derived from biological sources, such as
plants, animals, or microorganisms. They can be classified into several types as
follows:
• Natural Biomaterials: These are materials found in nature, such as wood, cotton,
silk, and collagen.
• Synthetic Biomaterials: These are materials that are chemically synthesized but
mimic the properties of natural biomaterials. Examples include biodegradable
polymers such as polylactic acid (PLA) and polyhydroxyalkanoates (PHAs).
• Hybrid Biomaterials: These are combinations of natural and synthetic biomateri-
als, often designed to have specific properties for targeted applications.
Biomaterials have significant applications in the medical and healthcare fields.
They are used in implantable medical devices, tissue engineering, drug delivery
systems, wound healing, and regenerative medicine. Examples include biocompat-
ible metals, ceramic implants, biodegradable sutures, and scaffolds for tissue regen-
eration. Biomaterials are increasingly being explored as alternatives to conventional
petroleum-based plastics in packaging. These biomaterials can be derived from
sources such as starch, cellulose, or polylactic acid (PLA). They offer the advantage
of being renewable, biodegradable, or compostable, reducing the environmental
impact of packaging waste.
Biomaterials are utilized in the textile industry to produce sustainable and eco-­
friendly fabrics. Fibers derived from natural sources such as bamboo, hemp, or soy-
bean, as well as bio-based polymers, are used to create biodegradable and renewable
textiles. Biomaterials are also finding applications in the construction industry. For
example, bioplastics made from plant-derived polymers can be used in building
materials such as insulation, coatings, and composites. Biomaterials can offer
advantages such as reduced carbon footprint, improved energy efficiency, and
enhanced sustainability in construction projects. Biomaterials contribute to the
reduction in greenhouse gas emissions and the conservation of non-renewable
resources. As they are derived from renewable biological sources, they have the
potential to replace fossil-based materials and reduce reliance on petroleum-derived
products. Biomaterials research continues to advance, focusing on developing new
materials with improved properties, functionality, and performance. This includes
exploring biofabrication techniques, nanomaterials, and smart biomaterials that
respond to environmental stimuli. Biomaterials represent a growing field within the
bioeconomy, offering sustainable alternatives to traditional materials. They provide
opportunities to reduce environmental impact, support innovative applications
Bioeconomy: A Sustainable Approach for Biomass Waste Management 247

across industries, and contribute to the transition toward a more sustainable and
circular economy.
The bioeconomy emphasizes the efficient use of biomass and waste streams,
promoting the circular economy. It includes the conversion of organic waste into
biogas or biofertilizers and the utilization of by-products from various industries for
value-added purposes. Waste management and the circular economy are closely
linked concepts within the bioeconomy framework. They both focus on minimizing
waste generation, promoting resource efficiency, and creating a closed-loop system
where materials are reused, recycled, or repurposed. The bioeconomy encourages
waste reduction and prevention strategies in various industries, including agricul-
ture, forestry, and manufacturing. This involves minimizing waste generation
through improved production processes, product design, and consumption patterns.
The circular economy emphasizes recycling and repurposing of materials to extend
their lifecycle. In the bioeconomy, this can involve recycling agricultural residues,
forest biomass, and other organic waste streams into valuable products such as com-
post, biofuels, bio-based chemicals, and biomaterials. The bioeconomy promotes
the efficient use of biomass resources, including organic waste, for bioenergy pro-
duction. Biomass can be converted into biogas through anaerobic digestion or used
in bioenergy systems to generate heat, electricity, and biofuels. This reduces the
reliance on fossil fuels and contributes to a more sustainable energy system.
The bioeconomy supports the use of sustainable packaging materials, such as
bio-based plastics and biodegradable materials, which can be derived from renew-
able resources and are designed to have a reduced environmental impact compared
to conventional packaging. By implementing effective waste management practices
and embracing the principles of the circular economy, the bioeconomy aims to min-
imize waste, maximize resource efficiency, and reduce environmental pollution.
These approaches contribute to a more sustainable and resilient economic system
that values the responsible use and management of resources throughout their life-
cycle. Advanced technologies, such as genetic engineering, bioprocessing, and syn-
thetic biology (Fig. 1), play a crucial role in enabling the production of bio-based
products and the development of sustainable processes within these industries
(Keshava et al., 2018). The bioeconomy encompasses a wide array of sectors and
technologies that leverage renewable biological resources to produce a diverse
range of products, fostering sustainable economic growth and environmental
preservation.
Biorefineries are key components of the circular bioeconomy. They are facilities
that process biomass into multiple valuable products, similar to how traditional
refineries process crude oil. Biorefineries utilize various technologies such as bio-
chemical and thermochemical processes to produce bio-based chemicals, materials,
and energy from biomass feedstocks. The bioeconomy encourages the use of com-
posting to recycle organic waste into nutrient-rich soil amendments. Compost can
be used to improve soil health and fertility in agriculture. Similarly, biofertilizers
derived from organic waste can provide essential nutrients to crops while reducing
the need for synthetic fertilizers. The circular economy promotes the development
of circular supply chains, where materials are reused, repaired, or recycled. In the
248 R. Mallick et al.

Biology

Systems
Engineering
Biology

Biology
Biotechnology Bioinformatics

Chemistry Mathematics

Fig. 1 Synthetic biology and its applications

bioeconomy, this can involve closed-loop systems for agricultural and forestry
products, ensuring that waste and by-products are reintegrated into the production
cycle rather than being discarded (Kardung et al., 2021).

2 Conclusion

The bioeconomy indeed offers a promising pathway toward a more sustainable and
circular economy. By shifting from a linear model of production and consumption
to a circular model, the bioeconomy aims to reduce waste, optimize resource use,
and promote the sustainable utilization of renewable biological resources. The bio-
economy emphasizes the efficient use of resources by valorizing biomass and waste
streams that were previously discarded or underutilized. By converting these
resources into bio-based products, energy, and materials, the bioeconomy mini-
mizes waste generation and maximizes resource efficiency. The bioeconomy
focuses on utilizing renewable biological resources, such as crops, forests, and
microorganisms, as the basis for production. These resources have the advantage of
Bioeconomy: A Sustainable Approach for Biomass Waste Management 249

being renewable, reducing dependence on finite fossil-based resources, and contrib-


uting to the transition to a low-carbon and sustainable economy. The bioeconomy
integrates circular practices by promoting the reuse, recycling, and repurposing of
materials. The bioeconomy plays a crucial role in integrating circular practices by
emphasizing the reuse, recycling, and repurposing of materials within the economy.
The bioeconomy encompasses the sustainable production and conversion of bio-
mass resources into bio-based products, including food, feed, biofuels, bioenergy,
and bio-based materials.
The bioeconomy focuses on utilizing renewable biomass resources, such as agri-
cultural residues, forestry by-products, and organic waste, as feedstock for various
applications. Instead of treating these materials as waste, they are repurposed and
transformed into value-added products, reducing reliance on non-renewable
resources. In the bioeconomy, bio-based materials are developed as alternatives to
traditional petroleum-based materials. These materials can be derived from agricul-
tural crops, forestry residues, or industrial by-products. By promoting the use of
bio-based materials, the bioeconomy reduces the consumption of fossil fuels and
contributes to the circular economy. The bioeconomy focuses on the production of
bioenergy and biofuels from biomass feedstock. This includes the conversion of
agricultural residues, energy crops, and organic waste into biofuels such as ethanol
and biodiesel. By utilizing these biofuels, which are derived from renewable sources,
the bioeconomy reduces greenhouse gas emissions and promotes a more sustainable
energy system.
Biorefineries are key components of the bioeconomy. They are facilities that
integrate various technologies to convert biomass into multiple products, including
fuels, chemicals, and materials. Biorefineries employ a cascading approach, where
different components of biomass are extracted and used for different purposes, max-
imizing resource efficiency, and minimizing waste. The bioeconomy encourages the
development of circular value chains, where the by-products and waste from one
process become feedstock for another. For example, in a bioethanol production pro-
cess, the residual biomass (stillage) can be used to produce biogas through anaero-
bic digestion, generating renewable energy and reducing waste. Bioeconomy seeks
to valorize waste materials by converting them into useful products. By utilizing
biotechnological processes, waste streams can be transformed into bio-based chem-
icals, fertilizers, and other valuable products. This approach reduces the environ-
mental impact of waste disposal and creates additional economic opportunities.
Integrating circular practices, the bioeconomy aims to maximize the efficient use of
resources, reduce waste generation, and contribute to a more sustainable and resil-
ient economy. It aligns with the principles of the circular economy by prioritizing
resource conservation, minimizing environmental impact, and promoting the reuse
and recycling of materials.
Biomass waste and by-products can be transformed into valuable products
through processes such as composting, biorefining, and bioconversion. This circular
approach minimizes waste sent to landfills, reduces environmental pollution, and
contributes to a more sustainable resource management system. The bioeconomy
offers environmental benefits by reducing greenhouse gas emissions, promoting
250 R. Mallick et al.

biodiversity conservation, and mitigating environmental degradation. Sustainable


farming practices, bioenergy production, and the use of bio-based materials contrib-
ute to lower carbon footprints and a more sustainable use of natural resources.
By capitalizing on the potential of bio-based products, the bioeconomy generates
employment opportunities across the value chain, from agriculture and manufactur-
ing to research and development. The bioeconomy stimulates innovation and
research by driving advancements in biotechnology, sustainable agriculture, bioen-
ergy, and biomaterials. These innovations enable the discovery of new processes,
products, and applications, driving continuous progress toward a more sustainable
and circular economy. By embracing the principles of the bioeconomy, we can shift
toward a more sustainable and circular economic model that optimizes resource use,
reduces waste, and contributes to environmental protection. This transition has the
potential to bring economic, social, and environmental benefits, paving the way for
a more sustainable future (Peppas & Langer, 1994).

References

Birner, R. (2018). Bioeconomy concepts. In Bioeconomy: Shaping the transition to a sustainable,


biobased economy (pp. 17–38). Springer.
Bugge, M. M., Hansen, T., & Klitkou, A. (2016). What is the bioeconomy? A review of the litera-
ture. Sustainability, 8(7), 691.
Carus, M., & Dammer, L. (2018). The circular bioeconomy – Concepts, opportunities, and limita-
tions. Industrial Biotechnology, 14(2), 83–91.
Dietz, T., Börner, J., Förster, J. J., & Von Braun, J. (2018). Governance of the bioeconomy: A
global comparative study of national bioeconomy strategies. Sustainability, 10(9), 3190.
El-Chichakli, B., von Braun, J., Lang, C., Barben, D., & Philp, J. (2016). Policy: Five cornerstones
of a global bioeconomy. Nature, 535(7611), 221–223.
Kardung, M., Cingiz, K., Costenoble, O., Delahaye, R., Heijman, W., Lovrić, M., van Leeuwen,
M., M’barek, R., van Meijl, H., Piotrowski, S., & Ronzon, T. (2021). Development of the cir-
cular bioeconomy: Drivers and indicators. Sustainability, 13(1), 413.
Keshava, R., Mitra, R., Gope, M. L., & Gope, R. (2018). Synthetic biology: Overview and applica-
tions. In Omics technologies and bio-engineering; towards improving quality of life, Volume 1,
Emergingfields, animal and medical biotechnologies (pp. 63–93). Academic/Elsevier.
McCormick, K., & Kautto, N. (2013). The bioeconomy in Europe: An overview. Sustainability,
5(6), 2589–2608.
Patermann, C., & Aguilar, A. (2018). The origins of the bioeconomy in the European Union. New
Biotechnology, 40, 20–24.
Peppas, N. A., & Langer, R. (1994). New challenges in biomaterials. Science, 263(5154),
1715–1720.
Pfau, S. F., Hagens, J. E., Dankbaar, B., & Smits, A. J. (2014). Visions of sustainability in bio-
economy research. Sustainability, 6(3), 1222–1249.
Stegmann, P., Londo, M., & Junginger, M. (2020). The circular bioeconomy: Its elements and
role in European bioeconomy clusters. Resources, Conservation and Recycling: X, 6, 100029.
Zilberman, D., Kim, E., Kirschner, S., Kaplan, S., & Reeves, J. (2013). Technology and the future
bioeconomy. Agricultural Economics, 44(s1), 95–102.
Application of Flower Wastes to Produce
Valuable Products

Avnish Chauhan, Manya Chauhan, Muneesh Sethi, Arvind Bodhe,


Anirudh Tomar, Shikha, and Nitesh Singh

1 Introduction

Managing the organic floral waste in the nation amounts to 4.74 × 106 t/d and is
thrown away after use in numerous religious rituals, social events, etc. (Srivastav
& Kumar, 2020). It is well known that most plant components, including floral
extract, are rich in phytochemicals. Flowers are an essential component of human
life, and people use them to decorate for key occasions such as weddings, birth-
day parties, and celebrations, among many others. Numerous flowers are used in
food preparation, beverages, salads, and baking (Kelley et al., 2001, 2002).
Flowers provide the raw materials for wine (dandelion and elderflower),

A. Chauhan (*)
Department of Environment Science, Graphic Era Hill University, Dehradun, Uttarakhand,
India
M. Chauhan
Department of Economics, Indraprastha College for Women, New Delhi, India
M. Sethi
Research and Development, COER University, Haridwar, Uttarakhand, India
A. Bodhe
School of Engineering and Technology, G H Raisoni University Saikheda, Dhoda Borgaon
Mal, MP, India
A. Tomar
Department of Physics, Graphic Era Hill University, Dehradun, Uttarakhand, India
Shikha
Department of Chemistry, KGC, Gurukula Kangri (Deemed to be University), Haridwar,
India
N. Singh
Department of Plant Pathology, Faculty of Agricultural Science, SGT University, Gurugram,
Haryana, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 251
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_14
252 A. Chauhan et al.

vegetables (such as squash blossoms and rapini), spices (such as saffron, kewra,
and clove), and squash (Rose, burans flower). Flowers are readily available, con-
tain many phytochemicals, and are a sustainable and environmentally friendly
resource. They are abundant in xanthones, coumarins, terpenoids, sterol, and flavo-
noids (anthocyanins and catechins), which can be precursors for producing
nanoparticles (NPs) (Bachheti et al., 2020). According to the National Horticulture
of India, 322 thousand hectares were devoted to floriculture production in India in
2020–21, with 2,152,000 tonnes of loose flowers and 8,28,000 tonnes of cut flow-
ers produced in India (APEDA, 2023).
Solid waste, also acknowledged as the third pollution after air and water pollu-
tion, is any substance produced by human activity that is typically discarded or
undesirable. By applying the right technologies, waste can be turned into usable
products because it is a valuable raw material (Chauhan, 2019). The waste pro-
duced by temples, where a range of goods, including incense sticks, garlands,
sweets, coconut shells, flowers, grass, plant leaves, strings, and edible and non-
edible fruits, are dedicated to Gods (Samadhiya et al., 2017). Various religious
rituals are performed in India by people from different religions, such as Hindus,
Muslims, Christians, and Sikhs. This ritual is conducted in temples, mosques,
tombs, Mazars, and churches. There are various famous places in India where
flowers are used in huge quantities, and these are Badrinath, Haridwar, Kedarnath,
Katra, Ajmer, Mumbai, Khandwa, Pune, Rameshwar, Ujjain, Shirdi, Tirupati,
Bhubaneswar, Varanasi, etc., but do not have sufficient flower waste dis-
posal policy.
According to Masure and Patil (2014), over 40% of the total production of
flowers in India and Sri Lanka goes to waste because they are unsold. Especially
in areas regarded as key tourist destinations, the placement of floral debris along
roadsides and in open spaces gives the area a dirty appearance and distorts its
reputation (Waghmode et al., 2018). Biodegradable and non-biodegradable mate-
rials are included in the trash taken from the temple, and floral debris is separated
because it is biodegradable. As flowers are offered to Gods in practically all
religions and then abandoned later, a significant amount of flower waste is pro-
duced in religious settings, including temples, churches, and dargahs (Yadav
et al., 2015).
India is a nation of festivals, with many events taking place all year round.
Nearly all religious locations receive flowers as offerings from followers, but after-
wards, they are rendered useless and considered waste (Yadav et al., 2015). Famous
Indian holidays such as Dussehra, EID, Deepawali, Navratri, and Shivratri are
observed, and copious amounts of flowers are produced on these occasions.
Because flowers offered to God are revered as sacred objects, discarding them
with other garbage created by hotels, markets, restaurants, etc., would be offensive
to religious beliefs. They are instead abandoned in open areas or tossed into aquatic
bodies (Barad & Upadhyay, 2016; Mahindrakar, 2018; Kumar et al., 2020).
Application of Flower Wastes to Produce Valuable Products 253

2 Flower Production Worldwide

According to a study, India produces about 4738 t of F.W. (flower waste) daily
(Sharma et al., 2018). Due to the lack of strict laws, there is currently no adequate
system for the segregation, collection, transport, and management of F.W. from
markets and places of worship (Dutta & Kumar, 2021).
The marketing, creation, and commercial production of cut flowers, loose flow-
ers, cut greens, seeds, bulbs, and landscape plants are all included in the specialised
field of horticulture known as floriculture (Fig. 1).
The worldwide market for cut flowers, which was estimated to be worth US$
30.70 billion in 2020 but is now likely to be US$43.80 billion by 2027, is to increase
at a compound annual growth rate (CAGR) of 5.20% from 2020 to 2027 despite the
COVID-19 situation. The market for cut flowers in the USA is anticipated to grow
from $9.21 billion to $12.77 billion by 2028. According to Research and Markets
(2022), it is anticipated to grow at a CAGR of 4.8% from 2021 to 2028. According
to Research and Markets (2022), Germany is expected to grow within Europe at a
rate of roughly 4.2%. The rose, carnation, lily, chrysanthemum, and gerbera are the
five main types of cut flowers (Markets, 2021). The Netherlands, one of the nations,
has a long record in the industry dating back to the early twentieth century. More
than 40% of all exports are currently cut flowers, making it the world’s largest
exporter (Garcia, 2018; Darras, 2020; Republic, 2021).

3 Flower Production in India (Fig. 2)

The traditional agricultural practice of floriculture in India has a great deal of prom-
ise to enable small and marginal farmers to launch successful side businesses.
Demand for floriculture goods has increased globally in both developed and

Country with the Highest Flower Production


(2018)

Netherlands
Columbia
Ecuador
Kenya
Belgium
Ethiopia
Malaysia
Italy

Fig. 1 Production of flowers in 2018 worldwide


254 A. Chauhan et al.

Flower Production in India (2022)

Kerala
Taminadu
Karnataka
Madhya Pradesh
Uttar Pradesh
Other Indian State

Fig. 2 Production of flowers in India (2022). (Source: APEDA (2023))

developing nations. As a result, it has flourished as a global and Indian agricultural


sector. Production and commerce in floriculture have constantly increased during
the last 10 years. The trade of flowers, the creation of nursery and potted plants, the
production of seeds and bulbs, micropropagation, and the extraction of essential oils
are all part of the floriculture industry in India. In 2020–21, India exported nearly
15,695.31 MT of floriculture products worth 77.84 USD Millions.
According to the Indian government, cut flowers, pot plants, cut foliage, seeds,
bulbs, and tubers make up the majority of floriculture products. They also include
rooted cuttings and dried flowers or leaves. The tulip, gerbera, gladiolus, orchid,
chrysanthemum, rose, and chrysanthemum are the most significant floricultural
crops in the world. Green buildings are used for the cultivation of crops used in
floristry, such as gerberas and carnations. The crops grown in open fields include
chrysanthemums, roses, gaillardia, lily marigolds, aster, and tuberose, among others.
Horticulture Crops for 2018–19 figures show that there were 303 thousand hect-
ares of flower crops in total. India only trails China in the amount of land that is
entirely devoted to floriculture. According to estimates, 647 thousand MT of the
2910 thousand MT of flowers produced were cut and loose flowers. India is the
world’s leading exporter of fresh and dried cut flowers. With a total production of
253.24 thousand tonnes, Karnataka is the first-place state among the states. With
roughly 53.26 thousand hectares of land cultivated for floriculture, Kerala is leading
in terms of area (Vikaspedia, 2023).

4 Components of the Floriculture Industry

1. Cut flower cultivation: Cut flower cultivation is highly common throughout the
nation. Examples of cut flowers are roses, gerberas, tuberoses, gladioli, and
chrysanthemums. Orissa, Karnataka, MH, AP, UP, West Bengal, and Gujarat
are the top states of India producing cut flowers.
Application of Flower Wastes to Produce Valuable Products 255

2. Loose flower cultivation: Due to the great domestic demand, loose flowers,
which include numerous traditional flowers, including marigolds, China asters,
jasmine, Cassandra, and bacteria, are grown extensively throughout the nation.
The top three states for producing loose flowers are Andhra Pradesh, Tamil
Nadu, and Karnataka.
3. Protected cultivation: In recent years, the area under protected cultivation has
increased from 500 ha to over 5000 hectares. Major flowers grown in poly
houses include roses, gerberas, and carnations, primarily in Gujarat,
Maharashtra, and Karnataka. Sikkim, Arunachal Pradesh, Goa, and Kerala all
have net houses or playhouses where orchids and anthurium are grown.
4. High-value flower crops: Due to their high floral value, high-value flowers,
including Asiatic ginger lilies, proteas, heliconias, orchids, and birds of para-
dise, are grown on significantly smaller plots of land than other flowers. These
expensive flowers are grown in the southern states of our nation.
5. Essential oils and floral scents: It is well known that flowers such as mogra,
lotus, champaca, nag champa, roses, jasmine, and tuberose are abundant in
essential oils. In India, these flowers have been used for thousands of years in
traditional medicine, perfumery, and aromatherapy, and they still play a signifi-
cant role in the culture and economy of the nation.
6. Production of flower seeds and bulbs: Punjab and Karnataka have a well-­
established seasonal flower seed production industry that is extremely profit-
able and caters to all three seasons. Plant production of bulbs, corms, and tubers
increased as human consumption did. Hybrid seed development is a lucrative
industry that ultimately aids in producing high-quality flowers.
7. Landscape gardening involves services such as landscape design and planning,
and construction and maintenance agreements. Due to the expanding popula-
tion, industries, and general awareness of environmental issues, this is a very
lucrative business. The landscaping of public spaces and societies, as well as
traffic island landscaping, is frequently demanded.
8. Allied industries related to floriculture: Allied industries related to floriculture
have emerged and are doing well due to high demand with the expanding area
under protected cultivation. Examples include greenhouse construction and
installation and shade net manufacturing.
9. Nursery: Growing ornamental plants for landscaping purposes is a lucrative
industry today. Nurseries specialising in annual plants, bulbous flowers, indoor
plants, outdoor plants, decorative plants, ornamental plants, commercial flow-
ering plants, shrubs, climbers, and tree seedlings are all doing well nationwide.
10. Floristry: The art of floristry entails creating and designing bouquets, arranging
flowers in both Western and Eastern designs, and decorating mandaps, cars,
banquet halls, etc., for various events such as performances, weddings, and
celebrations. This flower-based business generates large profits relatively
quickly and is quite lucrative.
11. Value addition: The drying or extraction of colour pigments can increase the
value of floriculture products by 30–100% for fresh flowers. About 70% of the
total floriculture export revenue comes from dry flowers. Unfortunately, despite
256 A. Chauhan et al.

this industry’s enormous export potential, little attention has been paid to it.
Our nation accounts for 10% of the market for dry flowers worldwide. India’s
top export markets are the USA, the UK, the Netherlands, Germany, and Italy.
Flowers and plant parts are typically gathered from wild sources to make dried
flowers. In India, the dried flower industry is primarily centred in Tuticorin
(Tamil Nadu) and Calcutta. Flowers can also be used to make potpourri, natural
dyes, and colours. Synthetic colours that are bad for human health are replaced
with natural dyes made from flowers such as marigold, hibiscus, Bixa, and
Butea monosperma (Sharma, 2019).

5 Future Prospects of India’s Floriculture

With numerous benefits, India has great potential to grow and prosper in this money-­
making industry. Below are a few of the benefits (Chawla et al., 2016):
(a) The country’s diverse agroclimatic conditions enable the year-round growth of
every type of flower in any region.
(b) India is strategically situated between two important export markets, namely
East Asia (Japan) and Europe (Holland), specifically between Wetland (near
Rotterdam) and Aalsmeer (near Amsterdam). Therefore, being close to these
important consumer markets makes for a great export platform.
(c) Markets in Japan, Russia, South-East Asia, and the Middle East are close to
India. The government permits cut flower and tissue-cultured plant exports to
receive air freight subsidies.
(d) From October to February, the winter season is moderate and excellent for
growing high-quality cut flowers without incurring additional costs for heating
greenhouses. Due to holidays such as Christmas, New Year’s Day, and
Valentine’s Day, demand is highest domestically and abroad. During this time,
European greenhouses—major producers—remain in a frozen state or require
heating.
(e) Unlike in Western nations, heating a greenhouse throughout the winter is
unnecessary in India. In the heat, cooling is not required either.
(f) Winter is when there are fewer holidays (Christmas, New Year’s, Valentine’s
Day, international events), and the major production hub of European nations
experiences frigid weather. Since flower production is decreasing in Europe
and costs are significantly higher in the winter, there is an increase in export
demand. European nations seek cheaper flowers from emerging production
hubs such as Asia (India, China) and Africa (Ecuador, Kenya) to offset their
higher production costs. India, endowed with various agroclimatic conditions,
can take the lead in producing the cut flowers that the world market demands.
(g) The availability of qualified labour.
(h) Rapidly growing domestic market.
Application of Flower Wastes to Produce Valuable Products 257

(i) The creation of multiple intense tissue culture units and growing awareness of
the need for quality production in nurseries.
(j) Ongoing studies on cutting-edge water and nutrition management in State
Agricultural Universities (SAUs) and research centres for cut flowers.
(k) The accessibility and adoption of high-tech irrigation and climate control
technologies.
(l) The accessibility and suitability of newly developed, adapted greenhouse tech-
nology for Indian circumstances.
(m) The accessibility of technical expertise in protected cultivation is appropriate
for our nation.
(n) Government policies that are pro-export and the construction of model floricul-
ture centres and a floriculture infrastructure park.

6 Generation and Disposal Methods of Flower Waste


in India

There may be several health problems if flower waste is improperly disposed of in


open landfills. After a few days of storage, microbes break down flower debris,
releasing poisonous gases into the atmosphere such as methane, carbon dioxide, and
ammonia that cause terrible odours and contribute to greenhouse gas emissions
(Singh et al., 2017). Aquatic ecosystems are at risk from the disposal of floral waste
in water bodies. Fish, diatoms, protozoans, and molluscs are among the aquatic
creatures that are significantly impacted by these waste disposal procedures
(Mahindrakar, 2018). While aquatic species become sick when insecticides and
synthetic fertilisers used for flower growth change the pH of water bodies.
Decomposed flowers contribute to extensive eutrophication by promoting algae
growth in water bodies. The removal of flowers may raise the organic load of a
water body, which may encourage the growth of noxious plants and microorganisms
and reduce oxygen levels (Makhania & Upadhyaya, 2015). Visitors also increase
along with the human population, which adds to the massive quantity of flower
debris produced (Samadhiya et al., 2017).
Temples are regarded as the residence of the gods. Hindus visit the temples to
receive blessings from their Gods as part of their culture before beginning any
momentous ceremony. Among them, individuals with a strong faith in God fre-
quently visit temples. Flowers are used during worship. As a result, flower waste
from temples around the world is enormous. Among the most common flowers
donated in temples are roses, marigolds, jasmine, and hibiscus. Managing the flower
waste produced by such operations has become an emerging problem because it
harms numerous life forms.
India’s diverse environment promotes the cultivation of a vast range of flower
species, frequently utilised as decorations in places of worship for special events. In
temples, various religious rituals are carried out, during which various offerings are
258 A. Chauhan et al.

made to the Gods, including sweets, leaves, garlands, fruits, and flowers (Samadhiya
et al., 2017).
Flowers are discarded from various locations, including hotels, weddings, gar-
dens, temples, churches, Dargahs, and other ceremonial settings. Religion is a way
of life in India. It is a fundamental component of Indian culture as a whole. People
regularly visit temples to present flowers, fruits, coconuts, sweets, and other offer-
ings as part of their adoration of God. In water bodies, most flowers, diverse plant
leaves, milk, and curd are subsequently discarded (Singh & Singh, 2007).
Worshippers in temples constantly present these flowers, so they need to be used
more and end up as waste. India is a country that observes numerous holidays all
year long, which inevitably leads to the creation of solid waste. It is important to
carefully examine this portion of trash because it is commonly ignored. Because of
our religious beliefs, many of us don’t throw away flowers and other items used for
prayers; instead, packed in plastic bags and throw them into rivers. Additionally,
flowers are stored beneath cherished trees, making proper disposal impossible. In
case of Banaras, one of the nation’s holiest cities, lacks a plan for getting rid of the
tons of trash that its various temples produce. The rubbish that is left behind in the
city of temples weighs between 3.5 and 4 tonnes every day (Mishra, 2013).
Compared to the degradation of kitchen waste, the degradation of floral waste is
a very time-consuming process (Jadhav et al., 2013). Consequently, a proper and
environmentally friendly method of treating floral waste is required. In certain stud-
ies, flower waste management and utilisation have been done. The Kashi Vishwanath
temple is one such instance, which receives the greatest number of devotees through-
out the year, particularly in the month of Shravan (beginning on July 23 and ending
on August 22). It has a system for getting rid of the hundreds of kilogrammes of
rubbish generated by devotee offerings; the floral waste produced in the temple is
turned into compost (Mishra, 2013).
Another instance of floral waste management that has produced positive results
is at Khwaja Moinuddin Chishti’s Ajmer Sharif Dargah, where daily flower offer-
ings of roughly 15–18 quintals were formerly disposed of in a well. The flowers are
recycled and give local women jobs. With technological help from the Central
Institute of Medicinal and Aromatic Factory, Lucknow (UP), the Ajmer Dargah
Committee created a rose water distillation factory in Ajmer City of Rajasthan
(Indian Express, May, 2010).
Every year, hazardous herbicides and insecticides are dumped into the river
along with more than eight million tonnes of flowers. These insecticides and pesti-
cides were used to cultivate the flowers (UNEP, 2019).
Most temple waste comprises organic materials such as flowers, coconut shells,
incense stick residue, fruits, and leaves, eventually making their way into trash cans
or rivers and causing pollution and hygiene issues. Thus, the numerous approaches
documented for utilising temple trash have been discussed in the present research.
Application of Flower Wastes to Produce Valuable Products 259

7 Composition of Flower Wastes

The Netherlands alone contributes over 33% of the entire production of flowers in
the European Union (EU), which accounts for roughly 44% of the global total
(Abeliotis et al., 2016). Sepals, petals, stamens with pollen-bearing anthers, an
ovary with a stigma and style, and nectaries are only a few of the components that
make up flowers (Miller et al., 2011).
According to Elango and Govindasamy (2018), the composition of floral waste
varies depending on the location. For example, in Dargahs, jasmine flowers make up
most of the trash, whereas, in Gurudwaras, marigold flowers predominate. In tem-
ples, marigold, lotus, rose, etc., are used.
Cut flowers, floral arrangements, the creation of essential oils, and other flower-­
based goods are only a few of the sources of flower wastes. Depending on the spe-
cies of flower, its stage of development or processing, and the particular waste
stream under consideration, the composition of floral wastes can change. However,
some common components of flower wastes include the following:
1. Organic matter: Flower wastes typically contain a significant amount of organic
matter, such as plant stems, leaves, and petals.
2. Water: Flowers and other plant materials contain a high percentage of water,
which contributes to the weight and volume of flower wastes.
3. Nutrients: Flowers contain a variety of nutrients, such as nitrogen, phosphorus,
and potassium, which can be beneficial for soil and plant health.
4. Pesticides: Flowers may be treated with pesticides or other chemicals during
production or transport, which can be present in the resulting waste stream.
5. Plastics: Some flower wastes may contain plastics or other non-biodegradable
materials from packaging or floral foam used in arrangements.
6. Other contaminants: Flower wastes may also contain other contaminants, such
as metals, dyes, or preservatives, depending on the specific waste stream being
considered.
Depending on the species of flower, the chemical composition can change. A
flower typically consists of a variety of organic substances, including lipids, organic
acids, proteins, carbohydrates, and amino acids. Additionally, flowers include the
carotenoids, flavonoids, and anthocyanin pigments that give them their distinctive
colours. Additionally, these pigments have antioxidant qualities. Essential oils,
which are composed of volatile organic substances, may also be found in flowers.
These oils contribute to the flower’s fragrance and might potentially be medicinal.
Overall, a flower’s chemical makeup is intricate and can change based on its species
and stage of growth. Flavonoids, volatiles, myricetin, and quercetin can all be found
naturally in chrysanthemum flowers (Wu et al., 2010). Jasmine flower, also known
as Jasminum, contains steroids, essential oils, alkaloids, glycosides, flavonoids,
phenolics, and saponins (Kunhachan et al., 2012).
According to Thakare et al. (2017), tannins, anthocyanins, riboflavin, organic
acid, sugars, tannins, pectin, carotenoids, mineral salts, and salt of tartaric acid are
260 A. Chauhan et al.

all abundant in rose blossoms. Essential oils, anthocyanin, tannins, quinines, alka-
loids, polysaccharides, protein, carbohydrates, reducing sugars, and steroids are
only a few of the many substances found in Hibiscus rosa-sinensis (Al-Snafi, 2018).
The majority of Tagetes species contain thiophenes, flavonoids, carotenoids, pheno-
lic chemicals, and terpenoids, according to research by Gupta and Vasudeva (2012).
Lotus flowers include a variety of alkaloids, flavonoids, and non-flavonoid chemi-
cals (Paudel & Panth, 2015).

8 Utilisation of Flower Waste

These are just a handful of the numerous applications for floral debris. We can cut
trash and make new, beautiful, and practical goods by recycling floral waste.
Moreover, the blooms can be used in herbal items such as herbal colours and natural
dyes. The most prevalent offering in temples, besides coconut shells, is flowers. The
proper disposal of flowers is a crucial problem for protecting the environment.
These discarded flowers can be turned into useful resources by using the proper
disposal techniques or strategies. Several start-ups have also emerged in recent
years, which are using innovative technologies to convert flower waste into products
such as organic dyes, handmade papers, and natural fragrance oils. Methods includ-
ing vermicomposting, composting, dye and essential oil extraction, making holi
colours, and biogas production are mentioned in the literature (Fig. 3).
1. Vermicomposting
Vermicomposting is a prominent technique to convert waste material into useful
materials without adversely affecting the environment. Although vermiculture has
been practised for at least a century, it is now accepted on a global scale for a variety

Fig. 3 Shows potential paths for using FW in several fields


Application of Flower Wastes to Produce Valuable Products 261

of ecological goals, including waste management, soil detoxification, regeneration,


and sustainable agriculture. Waste has accumulated more due to the expansion of
companies and the continuously growing human population (Joshi & Chauhan,
2006; Chauhan & Joshi, 2010). Waste non-utilisation is a problem that is reduced by
recycling wastes using Vermitechnology (Chauhan et al., 2010). Composting is a
method that can be utilised with flower waste. It is nutrient-rich, and when com-
bined with other organic materials such as grass clippings, leaves, and food wastes,
it can produce a compost that is high in nutrients and can be used to fertilise plants.
Worms were used to decompose the Nirmalya (temple garbage) that Gaurav and
Pathade (2011) collected from the Ganesh temple in Sangli, Maharashtra. The mix-
ture of cow manure, temple debris, and biogas digester effluent was left to disinte-
grate for 30 days at 30 °C. In pot cultures, five flowering plants were evaluated
while being fertilised with prepared vermicompost. In comparison to the control
sets that did not get the vermicompost treatment, good growth metrics were obtained
in height, flowering time, number of flowering times, and flower yield.
Vermicomposting flower waste is an excellent and ethical way to handle floral waste.
To handle floral waste, Shouche et al. (2011) used a variety of techniques, includ-
ing composting and vermicomposting. Several proportions of floral waste and cow
manure were utilised to create vermicompost. Numerous parameters, including
temperature, pH, and moisture content, which initially indicated some periodic
changes, were later reported to be constant.
According to Jadhav et al. (2013), a microbial consortium has been created to
decompose flower waste produced by temples efficiently. They extracted bacterial
cultures from soil samples near and around the temples. The obtained flower debris
was dried, combined with agar media, and then streaked with chosen soil samples
for isolation. The bio-manure consortium was determined to have good quality
without posing any environmental hazards, and the microbial consortium increased
the digestion of the waste.
Sailaja et al. (2013) worked on vermicomposting of flowers. Their investigation
also looked at prepared vermicompost’s nutritional level and microbial enumera-
tion. They concluded that plants cultivated in vermicompost grew faster than the
corresponding control. Auxin, gibberellins, and other plant hormones and enzymes
found in vermicompost stimulate plant development while discouraging plant dis-
eases. Vermicompost produced healthy plant yields as a result.
Tiwari (2014) conducted a study to use and manage floral debris collected from
10 well-known temples in Jaipur. Vermicomposting technology was applied to
lessen the amount of floral waste. Different ratios of marigolds (floral waste) were
collected, separated, and composted. The obtained vermicompost was assessed for
several factors: temperature, moisture content, organic carbon, accessible phospho-
rus, and pH. The study demonstrated that flowers make an excellent substrate for
vermicomposting.
The city of Tuticorin is home to a large number of dry flower processing and
exporting businesses. These industries generate a significant amount of organic
waste, most of which is floral. In 2014, Silvuai and Aneeshia tried to create useful
compost from this garbage. They used the fungi Pleurotus sapidus, Pleurotus
262 A. Chauhan et al.

flabellatus, and Ganoderma lucidum. It has been discovered that Pleurotus species
are highly effective for compost production and waste decomposition.
2. Flower Extraction of Essential Oils and Dyes
By using flower debris, natural dyes can be produced. You may produce a wide
range of hues by combining flowers and various mordents, which are compounds
used to fix the colour. Varied flowers produce varied colours. The best example of
turning flower waste into a useful approach is the extraction of colours and oils from
the leftover flower. Khan and Rehman (2005) evaluated a number of variables relat-
ing to the extraction and analysis of essential oils from Rosa species, including oil
yield, colour, and other physical and chemical features of two different kinds of
roses, Rosa damascene and Rosa centifolia. They came to the conclusion that the
chemical composition and scent components of the two species’ essential oils var-
ied both quantitatively and qualitatively.
A noteworthy quantity of flower waste is generated in Indian temples, according
to Vankar et al. (2009). This waste can be used to make commercial cotton, wool,
and silk dyeing dyes. They employed marigold flower petals, primarily composed
of carotenoids, lutein, and patulin, which have been found, extracted, and used as
dyes for textiles.
Perumal et al. (2012) surveyed about five temples in Chennai, Tamil Nadu, to
gauge the volume of flowers offered there. Every day, about 2.59 tons of flowers
were supplied, most of which were rose, marigolds, chrysanthemums, and jasmine.
They gathered rose petals from all kinds of flowers and shade-dried them to steam
distil the essential oils. Using the GC-MS technique, the chemical components of
rose oil were examined. There were 54 compounds identified, with phenyl ethyl
alcohol (23.19%) being the most prevalent. Other important compounds included
octadecane (10.49%), hexadecane (7.76%), phenyl ethyl decyl ester (5.77%), and
tetramethyl trisiloxane decanol (3.45%).
According to Ravishankar et al. (2014), over 13, 15,418 kg of flowers are offered
to the Gods in different national temples. The main flowers offered at Indian tem-
ples include rose, pinwheel flowers, jasmine, marigolds, guldavari, hyacinth, hibis-
cus, and tuberose. When these flowers are disposed of carelessly, disposal becomes
problematic. As a result, they published a study on using flower waste to extract
colours and essential oils. Flowers were ground up, dried, and then treated in etha-
nol, methanol, and hexane to extract the colours. Soxhlet equipment was utilised for
the extraction of essential oils. The fresh flower mixture was cooked and placed in
a Soxhlet apparatus with the appropriate solvent, and the resulting distillate pro-
duced the desired product. They claimed that the dye obtained in this way produces
good results and may be applied to clothing.
3. Biogas Generation
One of India’s 12 Jyotirlingas, Shri Mahakaleshwar Temple, generates about 3
tonnes of organic garbage yearly, most of which are flowers. Agarwal (2011) started
a study to create energy from temple garbage. He created vermicompost and biogas
Application of Flower Wastes to Produce Valuable Products 263

from floral and kitchen waste. Every 12 years, Mahakaleshwar hosts the Mahakumbh.
Thus, he tried creating a mechanism to produce electricity from that trash.
Singh and Bajpai (2012) investigated the anaerobic digestion of floral debris to
produce methane. To assess the purity of the gas produced, a gas chromatograph
was utilised, and the experiment was carried out in Lucknow’s chilly atmosphere.
They claimed that the procedure eradicated the negative environmental conse-
quences of flower disposal, eliminated pollutants such as biochemical oxygen
demand (BOD) and total solids (TS), and created biogas.
The waste of vegetables and flowers to produce biogas was studied by Ranjitha
et al. (2014). Cow dung was used as the inoculum for the study, conducted in a labo-
ratory anaerobic digester with a 1 L capacity. The findings demonstrated that flow-
ers produced more biogas (16.69 g/Kg) than vegetable waste (9.089 g/Kg), and their
digestion took less time. They also concluded that flowers, widely accessible in
India, make excellent feedstock for creating biogas, supporting the idea of turning
waste into wealth and improving sustainability.
4. Use of Flowers to Make Paper
Due to their role in advancing literacy and culture, paper products are vital to
society. Paper mills worldwide generate more than 400 million tonnes of paper
yearly. The three major paper-producing countries in the world—the USA, Japan,
and China—account for 50% of all paper output worldwide. The two biggest
importers and exporters of papers are Germany and the USA. Paper is still utilised
for many different things daily in the digital age and still greatly impacts society.
Paperboard and packaging paper are the most frequently made paper types, and
demand has recently increased due to the online shopping boom (Ogunwusi &
Ibrahim, 2014). Four hundred eight million tonnes of paper and paperboard were
used globally in 2021. By 2032, the anticipated rise in consumption over the follow-
ing 10 years will raise it to 476 million tonnes. Packaging accounts for the vast
majority of paper and paperboard output worldwide. Because more carbon dioxide
is emitted into the atmosphere due to using trees to make paper, this practice con-
tributes to global warming and deforestation (Albeiosu et al., 2019). Agriculture’s
production of trash affects the environment. The long-term issue can be avoided by
using novel technology, a scientific theoretical pathway, and food citation to high-
light the formation of a value-added product from an agricultural by-product.
Mahua flowers are readily available, inexpensive sources of carbohydrates that
can be found in non-agricultural settings such as a forest. Because the blossoms are
delicious, they are also used to make a variety of Indian culinary dishes, such as
kheer, puri, and halwa (Mishra & Pradhan, 2013). Mahua flowers have medicinal
value for several ailments and have aphrodisiac, galactagogues, and carminative
properties. Mahua flowers have high levels of reduced sugar, which makes up a
large portion of their total sugar content (Suryawansh & Mokat, 2020).
5. The Use of Flowers as Medicine
Pomegranate, or Punica granatum, is a species of deciduous tree of the
Punicaceae family. Pomegranate plant is grown in dry or semiarid environments and
264 A. Chauhan et al.

is now grown in most of the world, with the largest producers being India, Iran,
Turkey, China, and the USA. Punica granatum flower was shown to be both astrin-
gent and hemostatic. Its bloom has been linked to diabetes treatment in Unani and
Ayurvedic medical systems, while traditional Chinese medicine also uses it to heal
injuries, hair loss, and hair thinning. According to Sivarajan and Balachandran
(1994) and Wang et al. (2006), pomegranate flowers have been used medicinally in
the Ayurvedic, Unani, and Chinese medical systems.
The chemical makeup of tea flowers has also been observed to be identical to that
of the leaves and they contain significant amounts of total catechins (Su et al., 2000;
Lin et al., 2003). The tea flower extracts were also found to have antioxidant prop-
erty (Lin et al., 2003).
6. Removal of Heavy Metals
The biochar generated from floral waste has been prepared to extract copper
metal from synthetic wastewater. Additionally studied are its physicochemical char-
acteristics and adsorption mechanisms.
Biochar was produced using an acid–alkaline medium and strong carbonation.
The heavy metal (copper) wastewater was produced in a laboratory. In this study,
biochar was used in concentrations of 1%, 5%, 10%, 15%, 20%, and 25%. The
greatest removal efficiency and sorption capacity were 91.69% and 0.087 mmol/g,
respectively (Racha et al., 2022).
7. Biocidal Use
The flower species Tagetes sp. is a promising pest management option due to the
high quantity of thiophene discovered in the floral extract, which is thought to have
biocidal effects (Marotti et al., 2010). The red flower beetle, Tribolium castaneum,
which causes significant post-harvest damage to food grains, is resistant to the floral
extract of Tagetes erecta (Nikkon et al., 2009). After 96 hours of intake and topical
application, the insecticidal effects of methanolic extracts of Chrysanthemum sp.
flowers on Tribolium confusum exhibited a high mortality rate of up to 67% (Haouas
et al., 2008). After eating and topically applied, Chrysanthemum sp.’s larvicidal
effects on Spodoptera littoralis were also investigated by Haouas et al. (2010).
When fed with extract at concentrations between 1000 and 10,000 ppm, insect lar-
vae showed signs of toxicity in the form of increased mortality, slowed development
rates, and low weight gain.

9 Conclusion

This chapter addresses the issues associated with flower waste produced at religious
locations. The soil, water, and air quality of the surrounding ecosystem are seriously
impacted by improper flower waste treatment and disposal. But with the right pro-
cedures or processes, flower waste can be turned into a useful product. The thorough
analysis of numerous techniques for turning waste from temples into useful
Application of Flower Wastes to Produce Valuable Products 265

products, such as biogas, dyes, vermicompost, incense sticks, and concrete aggre-
gate replacement, suggests that the waste from temples can not only be disposed of
in a way that is both safe and environmentally friendly but can also be used to create
a variety of goods. This study will suggest an alternate method for managing gar-
bage since the waste will be used as a resource that will be recycled rather than
being burned or dumped in a landfill. It will shed light on how to cut down on the
waste produced by temples, eventually bringing in more money for them. Utilising
floral waste would eventually benefit society because it would allow people to live
in a neat and clean environment. The “green temple concept” may be useful in
developing government waste management policies and advocating a sustainable
development strategy for temples.

References

Abeliotis, K., Barla, S. A., Detsis, V., & Malindretos, G. (2016). Life cycle assessment of carnation
production in Greece. Journal of Cleaner Production, 112, 32e38. https://doi.org/10.1016/J.
JCLEPRO.2015.06.018
Agarwal, S. (2011). Report on demonstration of the renewable energy system at Shri Mahakaleshwar
Temple Complex in Ujjain (pp. 1–4). Ministry of New and Renewable Energy.
Agricultural and Processed Food Products Export Development Authority (APEDA). (2023).
Ministry of Commerce and Industry, Government of India. https://apeda.gov.in/aped-
awebsite/six_head_product/floriculture.htm#:~:text=More%20than%2050%25%20of%20
the,Tamil%20Nadu%20and%20Madhya%20Pradesh
Alebiosu, S. O., Akindiya, I. O., Mohammed Raji, A., & Adefidipe, O. A. (2019). Investigating
the physical and mechanical properties of broadleaf cattail (Typha Latifolia) and water hya-
cinth (Eichhornia Crassipes) paper blend. International Journal of Engineering Research and
Technology, 8, 119–122.
Al-Snafi, A.E. (2018). Chemical constituents, pharmacological effects and therapeutic importance
of Hibiscus rosa-sinensis- A review. IOSR Journal of Pharmacy, 8(7), 1–5.
Bachheti, R. K., Fikadu, A., Bachheti, A., & Husen, A. (2020). A review of biogenic fabrication
of nanomaterials from flower-based chemical compounds, characterisation and their various
applications. Saudi Journal of Biological Sciences, 27, 2551–2562.
Barad, G., & Upadhyay, A. (2016). Degradation of flower wastes: A review. International Journal
for Scientific Research & Development, 4(4), 1–2.
Chauhan, A. (2019). Environmental pollution and management (pp. 1–232). Wiley Publication
Publishing House.
Chauhan, A., & Joshi, P. C. (2010). Composting of some dangerous and toxic weeds using Eisenia
foetida. Journal of American Science, 6(3), 1–6.
Chauhan, A., Kumar, S., Singh, A. P., & Gupta, M. (2010). Vermicomposting of vegetable wastes
with cow dung using three earthworm species Eisenia foetida, Eudrilus eugeniae and Perionyx
excavatus. Nature and Science, 8(1), 33–43.
Chawla, S. L., Patil, S., Ahlawat, T. R., & Agnihotri, R. (2016). Present status, constraints and
future potential of floriculture in India. Commercial Horticulture, 1, 29–38.
Darras, A. I. (2020). Implementation of sustainable practices to ornamental plant cultivation world-
wide: A critical review. Agronomy, 10(10), 1570. https://doi.org/10.1016/j.jclepro.2020.124669
Dutta, S., & Kumar, M. S. (2021). Potential of value-added chemicals extracted from floral
waste: A review. Journal of Cleaner Production, 294, 126280. https://doi.org/10.1016/j.
jclepro.2021.126280
266 A. Chauhan et al.

Elango, G., & Govindasamy, R. (2018). Analysis and utilization of temple waste flowers in
Coimbatore District. Environmental Science and Pollution Research, 25(11): 1–7, https://doi.
org/10.1007/s11356-018-1259-0
Gupta, P., & Vasudeva, N. (2012). Marigold a potential ornamental plant drug. Hamdard Medicus,
55(1), 45–59.
Haouas, D., Flamini, G., Halima-Kamel, M. B., & Hamouda, M. H. B. (2010). Feeding perturba-
tion and toxic activity of five Chrysanthemum species crude extracts against Spodoptera lit-
toralis (Boisduval)(Lepidoptera; Noctuidae). Crop Protection, 29(9), 992–997.
Haouas, D., Halima-Kamel, M. B., & Hamouda, M. H. (2008). Insecticidal activity of flower and
leaf extracts from Chrysanthemum species against Tribolium confusum. Tunisian Journal of
Plant Protection, 3(2), 87–93.
Garcia, D. (2018). Cut-roses global value chain governance: Ecuadorian exports to the
Netherlands. The Hague University, MA thesis.
Gaurav, M. V., & Pathade, G. R. (2011). Production of vermicompost from temple waste
(Nirmalya): A case study. Universal Journal of Environment and Research Technology, 1(2),
182–192. https://doi.org/10.1007/s11356-­022-­239252
Indian Express. (2010, May). https://indianexpress.com/article/cities/lucknow/
with-­cimap-­help-­flowers-­at-­ajmer-­dargah-­to-­bring-­jobs/
Jadhav, A. R., Chitanand, M. P., & Shete, H. G. (2013). Flower waste degradation using microbial
consortium. Journal of Agriculture, 3(5), 1–4.
Joshi, P. C., & Chauhan, A. (2006). Composting of some organic materials using Eisenia foetida
and conventional microbial methods: A comparative study. Uttar Pradesh Journal of Zoology,
26(1), 123–125.
Kelley, K. M., Behe, B. K., Biernbaum, J. A., & Poff, K. L. (2001). Consumer preference for edible
flower colour, container size, and price. HortScience, 36, 801–804.
Kelley, K. M., Behe, B. K., Biernbaum, J. A., & Poff, K. L. (2002). Combinations of colours and
species of containerised edible flowers: Effect on consumer preferences. Horticultural Science,
37, 218–221.
Khan, M. A., & Rehman, S. U. (2005). Extraction and analysis of essential oil of rose species.
International Journal of Agriculture and Biology, 7(6), 973–974.
Kumar, V., Kumari, S., & Kumar, P. (2020). Management and sustainable energy produc-
tion using flower waste generated from temples. In V. Kumar, J. Singh, & P. Kumar (Eds.),
Environmental degradation: Causes and remediation strategies (Vol. 1, pp. 154–165). https://
doi.org/10.26832/aesa-­2020-­edcrs-­011
Kunhachan, P., Banchonglikitkul, C., Kajsongkram, T., Khayungarnnawee, A., & Wichet
Leelamanit, W. (2012). Chemical composition, toxicity and vasodilatation effect of the flow-
ers extract of Jasminum sambac. Hindawi Publishing Corporation, pp. 1–7, https://doi.
org/10.1155/2012/471312
Lin, Y. S., Wu, S. S., & Lin, J. K. (2003). Determination of tea polyphenols and caffeine in tea
flowers (Camellia sinensis) and their hydroxyl radical scavenging and nitric oxide suppressing
effects. Journal of Agricultural and Food Chemistry, 51, 975–978.
Mahindrakar, A. (2018). Floral waste utilisation. International Journal of Pure and Applied
Bioscience, 6(2), 1–5.
Makhania, M., & Upadhyay, A. (2015). Study of flower waste composting to generate organic
nutrients. International Journal of Innovative and Emerging Research in Engineering, 2(2),
145–147.
Markets, Ra. (2021). Cut flowers market forecast to 2028 - COVID-19 impact and global
analysis by flower type, application, and distribution channel. Research and Markets.
In this issue. https://www.researchandmarkets.com/reports/5178560/cut-­flower
s-­market-­forecast-­to-­2028-­covid-­19
Marotti, I., Marotti, M., Piccaglia, R., Nastri, A., Grandi, S., & Dinelli, G. (2010). Thiophene
occurrence in different Tagetes species: agricultural biomasses as sources of biocidal sub-
stances. Journal of the Science of Food and Agriculture, 90(7), 1210–1217.
Application of Flower Wastes to Produce Valuable Products 267

Masure, P. S., & Patil, P. B. M. (2014). Extraction of waste flowers. International Journal of
Engineering Research & Technology, 3, 43–44.
Miller, R., Owens, O. J., & Rørslett, B. (2011). Plants and colour: Flowers and pollination. Optics
and Laser Technology, 43(2), 282–294.
Mishra, N. (2013, Aug 17). Temple waste, a concern. Times of India. Retrieved from http://www.
timesofindia.indiatimes.com
Mishra, S., & Pradhan, S. (2013). Madhuca lonigfolia (Sapotaceae): A review of its traditional uses
and nutritional properties. International Journal of Humanities and Social Science Invention,
2(5), 30–36.
Nikkon, F., Habib, M. R., Karim, M. R., Ferdousi, Z., Rahman, M. M., & Haque, M. E. (2009).
Insecticidal activity of flower of Tagetes erecta L. against Tribolium castaneum (Herbst).
Research Journal of Agriculture and Biological Sciences, 5(5), 748–753.
Ogunwusi, A. A. and Ibrahim, H. D. (2014). Advances in pulp and paper technology and the impli-
cation for the paper industry in Nigeria. Industrial Engineering Letters, 4(10), 3–11.
Paudel, K. R., & Panth, N. (2015). Phytochemical profile and biological activity of Nelumbo
nucifera. Hindawi, pp. 1–16.
Perumal, K., Moorthy, T. A., & Savitha, J. S. (2012). Characterization of essential oil from offered
temple flowers Rosa damascena Mill. Asian Journal of Experimental Biology and Science,
3(2), 330–334.
Racha, U., Sukhavasi, P. C., Anuroop, P., & Chaitanya, R. (2022). Development of biochar from
floral waste for the removal of heavy metal (copper) from the synthetic wastewater. In IOP
Conference Series: Earth and Environmental Science (Vol. 1074, No. 1, p. 012033). IOP
Publishing.
Ranjitha, J., Vijayalakshmi, S., Vijaya, K. P., & Ralph, N. P. (2014). Production of bio-gas from
flowers and vegetable wastes using anaerobic digestion. International Journal of Research in
Engineering and Technology, 3(8), 279–283.
Ravishankar, R., Raju, A. B., Abdul, B. M., Mohapatra, A. K., & Kumar, M. (2014). Extraction of
useful products from temple flower wastes. Journal of Chemical Engineering and Research,
2(1), 231–239.
Republic, P. (2021). Floristry and floriculture industry statistics and trends. https://www.petal-
republic.com/floristry-­and-­floriculture-­statistics/#1-­global-­floricultur e-­production-­statistics.
Accessed 21 Dec 2021.
Research and Markets. (2022). https://www.prnewswire.com/news-­releases/cut-­flowers-­global-­
market-­r eport-­2 022-­f eaturing-­p rofiles-­o f-­f lorensis-­f lower-­s eeds-­u k-­o serian-­f lamingo-­
horticulture-­multiflora-­and-­washington-­bulb-­301500893.html
Sailaja, D., Srilakshmi, P., Shehanaaz, Priyanka, H., Bharathi, D. L., & Begum, A. (2013).
Preparation of Vermicompost from temple waste flowers. International Journal of Science
Innovation and Discoveries, 3(3), 367–375.
Samadhiya, H., Gupta, R. B., & Agrawal, O. P. (2017). Disposal and management of temple waste:
Current status and the possibility of vermicomposting. International Journal of Advanced
Research and Development, 2(4), 1–8.
Sharma, P. (2019). Floriculture-world wide production, trade, consumption pattern, mar-
ket opportunities and challenges. Medium. https://medium.com/@preetisharma_51610/
floriculture-­world-­wide-­production-­trade-­consumption-­pattern-­marke t-­opportunities-­and-­
challenges-­e797d87c87a7. Accessed: 2021-04-10.
Sharma, D., Yadav, K. D., & Kumar, S. (2018). Role of sawdust and cow dung on compost maturity
during rotary drum composting of floral waste. Bioresource Technology, 264, 285e289. https://
doi.org/10.1016/j.biortech.2018.05.091
Shouche, S., Pandey, A., & Bhati, P. (2011). Study about the changes in physical parameters during
vermicomposting of floral wastes. Journal of Environment and Research, 6(1), 63–68.
Singh, P., & Bajpai, U. (2012). Anaerobic digestion of flower waste for methane production: An
alternative energy source. Environmental Progress and Sustainable Energy, 31(4), 637–641.
268 A. Chauhan et al.

Singh, S. K., & Singh, R. S. (2007). Study on municipal solid waste and its management practices
in Dhanbad–Jharia coalfield. Indian Journal of Environmental Protection, 18(11), 850–852.
Singh, P., Borthakur, A., Singh, R., Awasthi, S., Pal, D. B., Srivastava, P., Tiwary, D., & Mishra,
P. K. (2017). Utilization of temple floral waste for extraction of valuable products: A close loop
approach towards environmental sustainability and waste management. Pollution, 3(1), 39–45.
Sivarajan, V. V., & Balachandran, I. (1994). Ayurvedic drugs and their plant sources. Oxford and
IBH Publishing Pvt. Ltd.
Srivastav, A. L., & Kumar, A. (2020). An endeavour to achieve sustainable development goals
through floral waste management: A short review. Journal of Cleaner Production. https://doi.
org/10.1016/j.jclepro.2020.124669
Su, S. K., Chen, S. L., Lin, X. Z., Hu, F. L., & Shao, M. (2000). The determination of ingredient of
tea (Camellia sinensis) pollen. Apicul China, 51, 3–5.
Suryawansh, Y. C., & Mokat, D. N. (2020). Variability studies in Madhuca longifolia var. latifolia
flowers from Northern Western Ghats of India. Indian Journal of Hill Farming, 33(2), 261–266.
Thakare, P.A., Deshbhratar, K. and Suryawanshi, M.N. (2017). A brief review on therapeutic
effects of – “ornamental plant” rose. International Journal of Ayurveda and Pharma Research,
5(12), 1–7.
Tiwari, P. (2014). Utilisation and management of floral waste generated in popular temples of
Jaipur City. The IIS University.
UNEP. (2019). A colourful solution to flower waste. https://www.unep.org/news-­and-­stories/story/
colourful-­solution-­flower-­waste.
Vankar, P. S., Sanker, R., & Wijayapala, S. (2009). Utilisation of temple waste flower- Tagetes erect
for dyeing of cotton, wool, silk on industrial scale. Journal of Textile and Apparel Technology
Manage, 6(1), 1–15.
Vikaspedia. (2023). https://vikaspedia.in/agriculture/farm-­based-­enterprises/floriculture
Waghmode, M. S., Gunjal, A. B., Nawani, N. N., & Patil, N. N. (2018). Management of floral
waste by conversion to value-added products and their other applications. Springer, 9(1), 1–11.
https://doi.org/10.1007/s12649-­016-­9763-­2
Wang, R., Wang, W., Wang, L., Liu, R., Ding, Y., & Du, L. (2006). Constituents of the flowers of
Punica granatum. Fitoterapia, 77, 534–537.
Wu, L. Y., Gao, H. Z., Wang, X. L., Ye, J. H., Lu, J. L., & Liang, Y. R. (2010). Analysis of chemical
composition of Chrysanthemum indicum flowers by GC/MS and HPLC. Journal of Medicinal
Plants Research, 4(5), 1–6.
Yadav, I., Juneja, S. K., & Chauhan, S. (2015). Temple waste utilisation and management.
International Journal of Engineering Technology Science and Research, 2, 1–6.
Myco-degradation of Lignocellulosic Waste
Biomass and Their Applications

Sahith Chepyala, Jagadeesh Bathula, and Sreedhar Bodiga

1 Introduction

Lignocellulose is a dry plant biomass that is prevalent on Earth and is regularly


synthesized and degraded in nature. A massive amount of lignocellulosic waste is
also generated by multiple industries such as forestry, pulp and paper, agriculture,
food from municipal solid waste, and animal wastes (Champagne, 2007; Wen et al.,
2004). This accounts for almost half of worldwide biomass production.
Lignocellulose accounts for 60% of the plant cell wall and is one of its key compo-
nents. Lignocellulose is a sugar biopolymer and its derivative is made up of lignin,
hemicelluloses, and cellulose (Galbe & Zacchi, 2012). Large amounts of lignocel-
lulose biomass are sometimes left unused and are either dumped or burnt carelessly,
causing major environmental problems by contributing to the emission of massive
amounts of carbon dioxide into the atmosphere. Lignocellulosic biomass contains
enormous amounts of energy that may be transformed into another form and used
for a variety of valuable products (Sánchez, 2009). Myco-degradation is one of
numerous environmentally acceptable methods for converting waste biomass into a
variety of value-added products. Basidiomycetes fungi can release a diverse variety
of wood-degrading enzymes that are involved in the breakdown of carbohydrate
components (cellulose and hemicellulose) as well as aromatic constituents (lignin)
(Kameshwar & Qin, 2016). These degradation-causing organisms employ lignocel-
lulose as a feedstock extremely for their development, reproduction, and subsequent
bioconversions (Chen & Chen, 2014). Multiple enzymes, including peroxidases,

S. Chepyala (*) · J. Bathula


Department of Forest Resource Management, Forest College and Research Institute,
Hyderabad, Telangana, India
S. Bodiga
Department of Basic & Social Sciences, Forest College and Research Institute,
Hyderabad, Telangana, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 269
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_15
270 S. Chepyala et al.

laccases, xylanases, endoglucanases, exoglucanases, β-xylosidase, and


β-glucosidases, work together to convert lignocellulosic complexes to monomers
(Hasunuma et al., 2013). As a result, these enzymes are critical for depolymeriza-
tion and degradation of lignocellulose into simple metabolites, and they are regarded
as the most efficient alternatives for the successful conversion of polysaccharide
into monosaccharide while avoiding the development of hazardous metabolites
(Pramanik & Sahu, 2017). Biotechnology based on lignin-degrading microbes and
their enzymes can help to more efficient and ecologically sound utilization of
renewable lignocellulosic feedstocks for sustainable production of materials, chem-
icals, biofuels, and energy in industrial applications. As a result, valorizing lignocel-
lulosic biomass using various microbial communities has been recognized as an
efficient and good strategy for a wide range of biotechnological applications.

2 Myco-degradation Mechanism and Enzymes Involved

The myco-degradation mechanism is highly reliant on the microbial communities’


degradative capacity and competence. The complex polymers of lignocelluloses are
broken down by a variety of microbes and potentially synergistic enzyme activity
(Parani & Eyini, 2010). Most fungi have two types of extracellular enzymatic sys-
tems: the hydrolytic system, which generates hydrolases responsible for polysac-
charide breakdown, and a distinct oxidative and extracellular ligninolytic system
that dissolves lignin and opens phenyl rings (Sánchez, 2009). Several microorgan-
isms, including fungi, can generate lignocellulolytic enzymes, which are involved
in the degradation of lignocellulosic biomass. The breakdown of hemicellulose, cel-
lulose, and lignin is known as hemicellulolysis, cellulolysis, and ligninolysis,
respectively.

2.1 Cellulolysis

Fungi have the potential to be the most powerful category of cellulose-degradable


microorganisms, producing a high amount of cellulases. Cellulolysis is the break-
down of cellulose, the major structural component of wood, by wood-decay fungi
into smaller glucose molecules that may be used as a carbon and energy source
(Chen & Chen, 2014). Fungi generate a variety of cellulolytic enzymes that degrades
cellulose. Endoglucanases, exoglucanases (cellobiohydrolases), and β-glucosidases
are the key cellulases involved in this process.
Endoglucanases initiate the cellulolytic process. These enzymes cleave the inter-
nal β-1,4-glycosidic linkages inside the cellulose chain, randomly breaking it down
into smaller oligosaccharides or soluble cellodextrins. Endoglucanases function at
numerous points throughout the cellulose chain, generating free ends for subse-
quent breakdown (Yamada et al., 2010). Exoglucanases breakdown the ends of cel-
lulose chains produced by endoglucanases. Cellobiose units (a dimer of glucose)
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 271

are released from the reducing and non-reducing ends of the cellulose chain.
Cellobiose units are basically pairs of glucose molecules connected by a β-1,4-­
glycosidic bond. Exoglucanases breakdown cellobiose units into individual glucose
molecules, which are then broken down further by β-glucosidases (Segato et al.,
2014). The cellobiose link is hydrolyzed by β-glucosidases, releasing two glucose
molecules. The glucose molecules that are released are subsequently transferred
into fungal cells and utilized as a carbon and energy source. Because of the syner-
gistic activity of the cellulolytic enzymes, the treatment is extremely efficient.
Endoglucanases provide free ends and accessible sites on the cellulose chain, allow-
ing exoglucanases to work more efficiently. Exoglucanases produce cellobiose,
which is easily hydrolyzed by β-glucosidases, boosting the total efficiency of cel-
lulose breakdown (Kuhad et al., 2011).

2.2 Hemicellulolysis

Endo-xylanase and β-xylosidase are involved in the hemicellulolysis mechanism in


wood-degrading fungi (Fig. 1), which is a two-step process that effectively breaks
down hemicellulose, particularly xylan, into its constituent sugar units. Hemicellulose
is a heterogeneous polymer made of multiple sugar units, and its breakdown neces-
sitates the production of a set of enzymes by the fungi (Van Gool et al., 2012).
Endo-xylanase is an enzyme that breaks down the internal β-1,4-glycosidic link-
ages in the xylan polymer, which is a key component of hemicellulose in wood. The
enzyme hydrolyzes the xylan chain at random sites, resulting in shorter xylo-­
oligosaccharides (xylo-oligos) of varying lengths. These xylo-oligosaccharides
generally include two to several dozen linked xylose units (Tani et al., 2014).

Lignocellulose

Cellulose Hemicellulose Lignin

Endo-
Endo-glucanase Li-peroxidase
xylanase

Exo-glucanase β-Xylosidase Mn-peroxidase

β-Glucosidase Laccase

Fig. 1 The three lignocellulose components (cellulose, hemicellulose, and lignin) and the cluster
of enzymes engaged in the hydrolysis of each component
272 S. Chepyala et al.

Endo-­xylanase adds additional reducing and non-reducing endings to xylan chains,


increasing the number of sites for future enzymatic activity. Following the action of
endo-xylanase, the resulting xylo-oligosaccharides are hydrolyzed further by
β-xylosidase. This enzyme cleaves the xylo-oligosaccharide’s terminal β-1,4-­
glycosidic link, releasing individual xylose molecules (monomers). Hemicellulose
(xylan) is successfully broken down into its monomeric form, xylose, by the com-
bined action of endo-xylanase and -xylosidase (Robl et al., 2013).

2.3 Ligninolysis

The breakdown of lignin, which has been widely explored using wood-rotting
organisms, particularly white-rot basidiomycetes fungus, is referred to as ligninoly-
sis (Wan & Li, 2012; Falade et al., 2017). Due to the resistant nature of lignin, ligi-
nolysis is a difficult process, but wood-degrading fungi have evolved specialized
enzyme systems that degrade this complex polymer. The refractory lignin is catabo-
lized into simpler metabolites by lignolytic enzymes during the ligninolysis pro-
cess. Though lignin degradation is carried out by several microorganisms, white-rot
fungus (WRF) have a greater potency than other organisms and are primarily
responsible for lignin breakdown (Tuomela et al., 2000).
Wood-degrading fungi generate a class of enzymes known as ligninolytic
enzymes. The principal lignin-degrading enzymes are lignin peroxidases (LiPs),
manganese peroxidases (MnPs), and laccases (Fig. 1). These enzymes are fre-
quently expressed when fungi are exposed to lignin or its breakdown products.
Ligninolysis is triggered by the activity of lignin peroxidases (LiPs). LiPs are flex-
ible enzymes that can oxidize a wide range of lignin-related compounds, including
non-phenolic and phenolic structures. They utilize hydrogen peroxide (H2O2) as a
co-substrate to catalyze the oxidative breakage of different chemical links in the
lignin polymer, resulting in free radicals (Wong, 2009). Manganese peroxidases
(MnPs) breakdown lignin together with LiPs. MnPs oxidize lignin using manganese
ions (Mn2+) and H2O2, producing free radicals that accelerate lignin breakdown
(Zhao et al., 2015). Laccases are copper-containing enzymes that contribute in the
oxidation of phenolic substances, including phenols generated from lignin. While
laccases cannot completely breakdown lignin, they do play a part in ligninolysis by
producing phenoxyl radicals. Laccases play a variety of critical roles in fungus,
such as participation in pigmentation (Clutterbuck, 1990), initiation of fruiting bod-
ies in mushrooms, sporulation, and effective lignin detoxification and degradation,
as well as pathogenicity (Mayer & Staples, 2002). The degradation of lignin pro-
duces a number of aromatic compounds as well as other lignin-derived chemicals.
Wood-degrading fungi have metabolic pathways that allow them to further catabo-
lize these chemicals and use the energy and carbon sources produced during lignin
breakdown.
Under both natural and controlled circumstances, several families of fungi may
release a wide range of lignocellulolytic enzymes. Fungi-secreted enzymes target
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 273

various links and connections in the chemical structure of lignocelluloses, simplify-


ing them into oligomeric metabolites. As a result, fungi in low-cost biodegradation
have built an appealing biological machinery for lignocellulosic material
bioconversion.

3 Fungi Used for Conversion

Microbial lignocellulose degradation is an essential stage in the carbon cycle


because it allows microbes to use plant sugars while overcoming the lignin barrier
(Martínez et al., 2005; Kersten & Cullen, 2007). Wood-degrading microorganisms
such as white-rot, brown-rot, and soft-rot fungi are used in biological lignin extrac-
tion to modify the chemical composition and structure of lignocellulosic biomass,
making it more accessible to enzyme digestion (Cardona & Sánchez, 2007).

3.1 White-Rot Fungi

White-rot fungi are a kind of wood-decaying fungi which can degrade lignin.
Approximately 90% of wood-rot fungi are white-rot fungi, which predominantly
attack lignin in wood. Unlike brown-rot fungi, which predominantly breakdown
cellulose in wood, white-rot fungi may degrade lignin, which is critical in biopulp-
ing. These fungi are classified into several taxonomic groupings, including
Basidiomycetes and Ascomycetes (Fragoeiro & Magan, 2005). Some well-known
white-rot fungi include species such as Phanerochaete chrysosporium, Trametes
versicolor, Polyporus berkeleyi, Pycnoporus cinnabarinus, Fome sulmarius,
Polyporus resinosus, Pleurotus ostreatus, and Ganoderma lucidum (Wan & Li,
2012). They are ubiquitous in woods and play an important role in the breakdown
of decaying wood. The first fungus utilized for pulp delignification under circum-
stances optimized for mineralization of synthetic lignin was Phanerochaete chryso-
sporium (Yu et al., 2009b).
While numerous fungi may breakdown wood cell wall polysaccharides, only
basidiomycetous white-rot fungi can degrade lignin effectively (Hatakka &
Hammel, 2011). This is because they have the unusual capacity to create extracel-
lular lignin-modifying peroxidases, which catalyze unspecific oxidative processes
to breakdown heterogeneous lignin structures. Furthermore, white-rot fungal
genomes contain many hydrogen peroxide-producing enzymes required for lignin-
modifying peroxidase activity (Lombard et al., 2014). Two white-rot fungal species
Obba rivulosa and Gelatoporia subvermispora provided all the lignocellulolytic
enzymes required for the white-rot form of degradation of solid spruce wood, which
included a variety of secreted cellulases, hemicellulases, lignin-modifying enzymes,
and auxiliary oxidoreductases (Marinovíc et al., 2022). White-rot fungi are also
known to be capable of degrading lignocellulosic waste biomass under heavy metal
274 S. Chepyala et al.

stress. Under lead stress, the activity of P. chrysosporium was examined, and it was
established that the fungus could be utilized as a reference for developing biotreat-
ment technology using P. chrysosporium to increase lignocellulosic waste biocon-
version and carbon cycle under contamination with heavy metals (Huang et al.,
2010). Lignin is the carbon source for white-rot fungi. White-rot fungi breakdown
lignin into carbon dioxide, allowing it to penetrate wood polysaccharides, which are
protected by lignin–carbohydrate complexes (Walter & Boyd-­Wilson, 2006).

3.2 Brown-Rot Fungi

Brown-rot fungi (BRF) are a kind of wood-decaying fungi that degrades cellulose
in wood while leaving lignin intact. These fungi are classified as Basidiomycota;
however, they lack lignin-degrading enzymes. BRF accounts for around 7% of all
known wood-decaying fungal species, with the majority thriving in conifers. Brown-­
rot fungi, in contrast to white-rot fungi, have acquired specialized mechanisms to
selectively breakdown cellulose while leaving lignin mostly intact (Eriksson et al.,
1990). In contrast to WRF, their decay is characterized by the decomposition of
carbohydrates (cellulose and hemicellulose) and substantial demethylation of lig-
nin, resulting in a brownish-appearing wood residue. The deterioration of brown-rot
fungi begins with the enzymatic breakdown of cellulose. These fungi create a class
of enzymes known as “cellulases,” which particularly target and degrade the cellu-
lose molecules in wood. Brown-rot fungi such as Postia placenta, Laetiporus por-
tentosus, Piptoporus betulinus, Serpula lacrymans, Sulphur shelf, and Gloeophyllum
trabeum can degrade wood polysaccharides but not oxidized lignin (Renvall, 1995).
Brown-rotted wood is dark and shrinking, and it is usually split into brick-shaped
or cubical fragments that readily decompose into brown powder (Blanchette, 1995).
Modified lignin is what gives wood its distinctive brown color. During brown rot,
the hemicelluloses in wood decay the fastest, followed by the removal of nearly all
the cellulose, leaving behind a complex, aromatic ring-containing polymer gener-
ated from the original lignin. Brown-rot fungi share certain degradative capacities
and routes with white-rot fungi. Both wood degradation pathways rely on radical
generation, low pH, and organic acid production. They raise the alkali solubility of
lignin and accelerate decay, all of which imply a critical role for radicals, particu-
larly in the early phases of decay (Kirk, 1975). Overall, their distinct cellulose-­
degrading abilities show promise for a variety of biotechnological applications.

3.3 Soft-Rot Fungi

Soft-rot fungi are a category of wood-decaying fungi that induce soft rot, a kind of
wood degradation. Soft-rot fungi breakdown both cellulose and hemicellulose in
wood, unlike white- and brown-rot fungi, which preferentially breakdown lignin
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 275

Table 1 Several recent studies with various fungal species for the breakdown of varied
lignocellulosic substrates and enzymes reported
Lignocellulose
SN Fungal species substrate Enzymes produced References
1 Pleurotus eryngii Peach waste Laccase Akpinar and Ozturk
Urek (2017)
2 Marasmius sp. Rice bran Laccase Vantamuri and
Kaliwal (2016)
3 Ganoderma lucidum Pineapple leaf LiP, MnP, and laccase Padma and Sudha
(2013)
4 Ganoderma Oak sawdust MnP and laccase Ćilerdžić et al.
applanatum (2016)
5 Phellinus robustus Food industry Laccase and MnP Songulashvili et al.
Ganoderma wastes (2006)
adspersum
6 Pleurotus ostreatus Banana waste Laccase, LiP, xylanase, Reddy et al. (2003)
and P. sajor-caju endo-glucanase, and
exo-glucanase

and cellulose, respectively. Soft rot is distinguished by wood softening and discol-
oration. These fungi are classified into two taxonomic groups: Ascomycetes and
Deuteromycetes. They may be found in a variety of settings, including soil, fresh-
water, and marine ecosystems (Goodell et al., 2008). Soft rot can develop in both
living trees (producing wetwood or slime flux) and processed wood products like
lumber and paper. Soft-rot fungal degradation includes the production of a cocktail
of enzymes that attack cellulose and hemicellulose. Cellulases, hemicellulases, and
pectinases are examples of these enzymes. Cellulases and hemicellulases convert
cellulose and hemicellulose polymers into simpler sugars, whereas pectinases
degrade pectin, which is a component of the middle lamella that keeps plant cells
together (Hamed, 2013). Soft rot has severe economic and environmental conse-
quences. Soft-rot decay can cause degradation of timber and wood products in the
forestry sector, diminishing their strength and longevity. It is a prevalent source of
wood deterioration in houses, utility poles, and other wooden structures. Soft-rot
fungi play an important role in the breakdown of woody waste in natural environ-
ments, contributing to nutrient cycle and carbon turnover (Schwarze et al., 2000)
(Table 1).

4 Applications in Varied Sectors

Agricultural leftovers (straws, hulls, stems, and stalks), deciduous and coniferous
woods, municipal solid wastes (MSW, paper, cardboard, yard garbage, and wood
products), waste from the pulp and paper industry, and herbaceous energy crops are
all examples of lignocellulosic biomass. These wastes are produced in huge
276 S. Chepyala et al.

quantities across the world and represent an environmental risk due to disposal
issues. To get rid of this bulk, these wastes are typically burned or left to decompose
in the field, resulting in unsanitary conditions and resource loss. The combustion of
these wastes not only leads in an oxygen-deficient atmosphere, but it also causes a
variety of respiratory disorders such as asthma, TB, and others, as well as impaired
sight at night (Agarwal et al., 2016). For hydrolytic enzymes like cellulases and
hemicellulases, which oversee degrading cellulose and hemicelluloses in the sub-
strate, respectively, the lignin polymers in these wastes function as a physical and
chemical barrier, making it challenging for them to access their substrates. As a
result, these wastes are challenging to breakdown and ruminants only break them
down slowly (Chaturvedi et al., 2018). The problem of the safe disposal of lignocel-
lulosic wastes and their resistance to breakdown can be handled by employing
fungi, namely the white-rot fungus (WRF), to degrade these wastes. With the help
of a potent enzymatic system, WRF can transform these wastes into products that
have added value, including nutritious food, enhanced animal feed, plant fertilizer,
biofuels, chemicals, and a low-cost supply of carbon for fermentation (Kuhad et al.,
2007). Over the past few years, lignocellulose biotechnology processes have
advanced significantly. For lignocellulosics, several new markets have developed
recently. The most ambitious of them has been the conversion of lignocellulose to
alternative energy sources, like gasoline ethanol (Sakamoto et al., 2012; Shahsavarani
et al., 2013). The pulp and paper sector has found that lignocellulose biotechnology
can increase process effectiveness, resulting in cost and energy savings. The key to
lignocellulose’s effective implementation in biotechnological projects will be elimi-
nating the lignin barriers that limit commercial exploitation of it (Asgher &
Iqbal, 2013).

4.1 Biofuels

In order to combat global warming, it is necessary to investigate alternate, more


affordable, and environmentally benign biofuel sources. This is due to the rising
expense of fossil fuels and their consequences on greenhouse gas emissions. One
possible strategy to boost ethanol output and satisfy global demand has been the
bioconversion of lignocellulosic biomass into fermentable sugars for the manufac-
ture of bioethanol (Kuila & Sharma, 2017). Lignocellulosic biomass is refractory,
which means that due to its complex structure, it is difficult to decompose and turn
directly into biofuels. Pretreatment is a crucial step that aims to alter the biomass
lignocellulosic structure in order to make it more accessible to subsequent biocon-
version procedures. Such biomass contains carbohydrates, which must be broken
down into simple sugars before being fermented to produce ethanol, a potential fuel
for vehicles (Balat & Balat, 2009). According to research, ethanol can be made
using sugar cane bagasse, corn cobs, coconut husks (copra), groundnut shells, other
nut shells, sawdust, cereal hay, corn stover, rice straw, and rice husks as raw
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 277

materials (Shahsavarani et al., 2013). For this transformation of biological resources,


such as crops high in energy or forestry residues, pretreatment of the feedstock is
necessary to hasten the enzymatic conversion of the feedstock into sugar, or sac-
charification. Pretreatment is used to extract sugars from these lignocellulosic
biomass-­based components, with glucose being the most important one for ethanol
production (Cardona & Sánchez, 2007).
There have been numerous reports on pretreatment techniques, including bio-
logical, chemical, physical, thermal, and enzymatic methods (Hamzeh et al., 2013;
Rohowsky et al., 2013). Given that lignin and hemicellulose in lignocellulosic mate-
rials are degraded by fungi like brown-rot, white-rot, and soft-rot, biological pre-
treatment is probably the most economical and environmentally friendly pretreatment
approach for those materials. Since no chemicals are used in this procedure, recy-
cling of chemicals is not required; therefore, there is no environmental effect
(Sindhu et al., 2016). In order to produce ethanol, García-Torreiro et al. (2016)
investigated the biological pretreatment of corn stover and examined the results of
pretreatment with the basidiomycete white-rot fungus Irpex lacteus. The fungus
was grown, and the feedstock was effectively delignified. The biomass after the
fungal pretreatment showed decreased lignin concentration, decreased sugar con-
sumption, and considerably increased total holocellulose digestibility, indicating
that the fungal pretreatment was beneficial in producing bioethanol by reducing
numerous operational expenses. According to Yu et al. (2009a), enzymatic hydroly-
sis of 17-week fungal-treated Chinese willow and China fir produced around 35%
and 18% of the reducing sugar yield, respectively.

4.2 Production of Enzymes

A key component of the contemporary biotechnology sector, enzyme production is


a rising area of biotechnology. Utilizing the potential of lignocellulosic waste mate-
rials is one of the most suitable methods to create affordable and effective enzymes
for biotechnological application reasons. Some of these mixes might contain high
levels of soluble carbohydrates and inducers of enzyme synthesis to produce ligni-
nolytic enzymes effectively (Asgher & Iqbal, 2013). Fungi are perfect candidates
for biotechnological applications because they have developed a variety of enzy-
matic machinery to breakdown lignocellulosic materials. This includes the manu-
facture of enzymes for numerous industrial processes. By utilizing the various
lignocellulosic substrates, these fungi have the extraordinary capacity to create
lignin-­degrading enzymes such as lignin peroxidases, manganese peroxidases, and
laccases, which play a significant role in many biotechnological applications. Due
to their assistance in removing the undesirable component of paper manufacturing
known as lignin, these enzymes have a tremendous impact on the pulp and paper
sector. Through synergistic interaction, mixed fungal cultures can increase the pro-
duction of enzymes; however, the outcome appears to depend on the specific species
278 S. Chepyala et al.

combination or mode of interaction between species and the microenvironmental or


nutritional conditions in the substrate under colonization (Chi et al., 2007).
According to Marđetko et al. (2021), the seven white-rot fungal species were used
to produce enzymes from corn cobs. Different productions of cellulolytic and ligno-
lytic enzymes have been observed. One good application of lignocellulolytic
enzymes (e.g., cellulases, xylanases, or laccases) is in food production.
Lignocellulolytic enzymes are frequently employed in the food industries for the
processing of fruits and vegetables, the texturizing and flavoring of food products,
the processing of juice, the extraction of oil, and for the fermentation of alcoholic
beverages. The widespread usage of these enzymes in many food processing sectors
speeds up food production and enhances the quality of the finished goods (Motta
et al., 2013; Van Dyk & Pletschke, 2012).

4.3 Composites

A potential answer to the sustainability and environmental problems that industries


and humanity confront is the use of nature-based resources, and more specifically,
biomass, instead of products made from petroleum. Recent years have seen a sig-
nificant increase in research into the use of lignocellulosic waste residual materials
in place of synthetic materials to create a variety of composites with diverse func-
tionality. Due to this, the creation of bio-based composites has drawn interest in the
field of materials science from both an ecological and environmental standpoint
(Bajpai et al., 2013). Because they are created from organic, regenerative, and bio-
degradable materials, lignocellulosic and fungal mycelium bio-composites are both
prospective replacements for petroleum-derived goods in the composites industry
(Hemmilä et al., 2017). Lignocellulosic bio-composites are made from wood or
agricultural waste that has been glued together with a type of adhesive. Traditionally,
these materials have been made into panels. Adhesives are typically applied to fur-
nish and are cured under heat and pressure to bond reconstituted lignocellulosic
biomass. The main adhesives used in the production of lignocellulosic bio-­
composites are formaldehyde and isocyanate-based synthetic resins, which are not
usually bio-based and raise issues for the environment and human health (Mantanis
et al., 2018; Solt et al., 2019).
Mycelium-based bio-composites, or goods where materials are bound together
by fungal mycelium, are a relatively new idea, but they have drawn significant sci-
entific and commercial attention (Jones et al., 2020). Typically, their production
begins with a low-cost lignocellulosic biomass substrate. The mycelia of filamen-
tous fungus are then put to a substrate and nutrient mixture in a mold with a speci-
fied form, where the mycelium partially digests and clings to the surface of the
substrate particles. The substrate particles are joined together into a single aggrega-
tion by the mycelia as they develop into entangled networks. Oven drying is fre-
quently used in both research and industrial labs to stop fungal development while
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 279

producing the finished product (Elsacker et al., 2020). In the case of hot-pressed
products, where mycelium development and enzyme secretion during fungal growth
are the same as the fungal or enzymatic pretreatment in lignocellulosic bio-­
composite manufacturing, it is reasonable to classify them in the same category as
conventional lignocellulosic bio-composites.

4.4 Fine Chemicals

The use of fossil fuels in the industrial manufacturing of a wide variety of chemicals
and synthetic polymers results in alarming environmental consequences including
global warming and littering issues, which have begun to pose a danger to the future
of the chemical and polymer industry (Serrano-Ruiz et al., 2011). One of the most
viable options that has been identified is biomass and products made from biomass.
These materials are produced by biological photosynthesis from accessible atmo-
spheric CO2, water, and sunlight (Zhou et al., 2008). As a result, biomass has been
regarded as the source of organic carbon that is sustainable on Earth and the ideal
substitute for petroleum in the manufacturing of fuels and fine chemicals that emit
no net carbon dioxide (Ragauskas et al., 2006). The most abundant and renewable
biomass on earth, lignocellulosic biomass, is crucial in this situation. Numerous
studies have demonstrated the immense potential of lignocellulosic biomass for the
generation of chemicals and fuels in a sustainable manner. To make cellulose and
hemicellulose more accessible and biodegradable for enzymatic or chemical action,
lignocellulosic biomass must be pretreated before it can be subjected to other types
of treatment in the manufacturing of chemicals (Barakat et al., 2013). Biotreated
lignocellulosic biomass has tremendous potential for the improvement of efficiency
and decreasing the cost of production. The pretreatment of biomass with WRF will
assist in reducing the biomass size and open its physical structure. The hydrolysis
process results in the production of a variety of products, including glucose (mostly
from cellulose and hemicellulose), xylose, mannose, galactose, and acetic acid
(from hemicellulose), and phenolic compounds (from lignin). One of the most valu-
able compounds made from agro-residual lignocellulosic biomass has been named
as xylitol (Kumar et al., 2008). It is an appealing choice for low-cost manufacturing
utilizing lignocellulosic materials because of its proven commercial uses in the food
and pharmaceutical sectors (Misra et al., 2013). A major source of xylose for micro-
bial fungal fermentation to produce xylitol is found in the hemicellulosic compo-
nent of waste materials, such as sugar industry wastes (sugar cane bagasse) (Chandel
et al., 2012). The selection of naturally occurring xylose-fermenting microorgan-
isms, such as fungi, yeasts, and bacteria, that can produce xylitol from agro-­
industrial lignocellulosic wastes, is emerging as a potential technique in this area
(Kamat et al., 2013; Misra et al., 2013).
280 S. Chepyala et al.

4.5 Pulp and Paper

In our everyday lives, paper is a product that is widely utilized for a variety of pur-
poses. Huge amounts of lignocellulosic materials are processed by the pulp and
paper industry. The industry employs leftover resources from various industrial pro-
cesses (such as sawmill wood chips, sugarcane processing bagasse) or fibers recov-
ered from recycled paper or paperboard in addition to directly collecting wood,
straw, bamboo, and other natural materials (Bajpai & Bajpai, 2018). There is a
growing market for paper, and advances in pulp processing technology have made
it possible to produce pulp and paper from a variety of lignocellulosic materials.
There are three main processing steps: (i) pulping, (ii) bleaching, and (iii) paper
manufacture (Singh et al., 2013). The biomass may be processed using conventional
pulping techniques including chemical and mechanical pulping to get rid of the
lignin. These techniques emit greenhouse gases and other dangerous pollutants into
the environment while harsh chemicals and high temperatures are used to separate
the fibers in wood. As an alternative, biological pretreatment of wood provides a
cost-effective, low-chemical-demanding, and environmentally friendly method to
address the difficulties involved in the production of pulp and paper. This process is
known as biopulping or myco-degradation (Kumar et al., 2020). According to
Masarin et al. (2009), Phanerochaete chrysosporium treatment of wood chips led to
successful biopulping of fresh E. grandis wood chips. Biotreated wood fibers sig-
nificantly reduced energy use and strengthened pulp.
Other by-products including fertilizers, soil improvements, and animal feed can
be produced from lignocellulosic biomass. The white-rot fungus breaks down lignin
and increase the biomass in vivo digestibility of dry matter. In an anaerobic environ-
ment, like the rumen, this improved digestibility produces organic carbon that may
be digested to organic acids (Saritha et al., 2012). Additionally, this waste may be
utilized to make biochar, which is added to soil to enhance its chemical and physical
properties (Bonga et al., 2020) (Fig. 2).

5 Optimization of Myco-degradation Conditions

Despite the myco-degradation having several benefits, it must be optimized in sev-


eral phases in order to improve the quality of the desired result. The capacity of vari-
ous fungal strains to breakdown the biomass varies. The choice of highly effective
fungal strains for breakdown is crucial. It is crucial to find and create new fungi that
are extremely active in every substrate being utilized since the fungal activity will
vary depending on the substrate. As none of the strains, including the most advanced
mutants, are capable of simultaneously producing high levels of enzymes required
for bioconverting lignocelluloses, it is necessary to identify consortia of effective
cellulolytic, lignolytic, or co-culturing microorganisms that may be able to syner-
gistically act to rapidly bioconvert agricultural residues without the need for
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 281

Biofuels

Pulp & Production


Paper of Enzymes

Lignocellulose

Fine
Composites
chemicals

Fig. 2 Numerous uses of lignocellulose in varied sectors

chemical pretreatment (Wang et al., 2022). In order to further improve the degrada-
tion of biomass, it is recommended to combine multiple fungal strains in co-cultiva-
tion. This can be especially beneficial when working with strains that have limited
enzyme activity for lignin degradation. Additionally, the addition of commercially
available or genetically modified fungal lignolytic enzymes can further enhance the
degradation of biomass. The primary obstacle to the commercialization of biofuel
and other derived products is the high cost of enzymes, which necessitates a 10- to
100-fold decrease in enzyme costs (Binod et al., 2019). For this reason, researchers
are continually tackling the challenges to reduce the cost of manufacturing lignocel-
lulose degradation enzymes by looking for novel producer organisms and manipu-
lating organisms or enzymes using modern techniques such as genetic engineering
(GE), metabolic engineering (ME), and strain enhancement.

6 Conclusion

The results of several studies on myco-degradation indicate that fungi produce


highly effective enzymes in the degradation and bioconversion of lignocellulose.
Various groups of fungi can secrete a variety of lignolytic enzymes in both natural
and man-made environments. The enzymes released by fungi attack various chemi-
cal bonds and linkages in the lignocellulose structure and simplify them to oligo-
meric molecules. As a result, the low-cost biodegradation of fungi has created an
attractive biological mechanism for bioconversion of lignocellulose substances.
Combining this type of biodegradation with the capitalization of natural
282 S. Chepyala et al.

lignocellulose residues into a low-cost, environmentally friendly, high-bearing


enzyme processing process results in the sustainable bioconversion of these sub-
stances into numerous byproducts of added value.

References

Agarwal, S., Vaseem, H., Kushwaha, A., Gupta, K. K., Maurya, S., Chaturvedi, V. K., Ravi, P., &
Singh, M. P. (2016). Yield, biological efficiency and nutritional value of Pleurotus sajor-caju
cultivated on floral and agro-waste. Cellular and Molecular Biology, 62, 1–5.
Akpinar, M., & Ozturk Urek, R. (2017). Induction of fungal laccase production under solid state
bioprocessing of new agroindustrial waste and its application on dye decolorization. 3 Biotech,
7, 1–10.
Asgher, M., & Iqbal, H. M. N. (2013). Enhanced catalytic features of sol–gel immobilized MnP
isolated from solid state culture of Pleurotus ostreatus IBL-02. Chinese Chemical Letters,
24(4), 344–346.
Bajpai, P., & Bajpai, P. (2018). Brief description of the pulp and papermaking process. In
Biotechnology for pulp and paper processing (pp. 9–26). Springer.
Bajpai, P. K., Singh, I., & Madaan, J. (2013). Tribological behavior of natural fiber reinforced PLA
composites. Wear, 297(1–2), 829–840.
Balat, M., & Balat, H. (2009). Recent trends in global production and utilization of bio-ethanol
fuel. Applied Energy, 86(11), 2273–2282.
Barakat, A., de Vries, H., & Rouau, X. (2013). Dry fractionation process as an important step
in current and future lignocellulose biorefineries: A review. Bioresource Technology, 134,
362–373.
Binod, P., Gnansounou, E., Sindhu, R., & Pandey, A. (2019). Enzymes for second generation
biofuels: Recent developments and future perspectives. Bioresource Technology Reports, 5,
317–325.
Blanchette, R. A. (1995). Degradation of the lignocellulose complex in wood. Canadian Journal
of Botany, 73(S1), 999–1010.
Bonga, C. P., Lima, L. Y., Leea, C. T., Ongb, P. Y., Jaromír, J., Klemešc, C. L., & Gaod, Y. (2020).
Lignocellulosic biomass and food waste for biochar production and application: A review.
Chemical Engineer, 81, 427–432.
Cardona, C. A., & Sánchez, Ó. J. (2007). Fuel ethanol production: Process design trends and inte-
gration opportunities. Bioresource Technology, 98(12), 2415–2457.
Champagne, P. (2007). Feasibility of producing bio-ethanol from waste residues: A Canadian
perspective: Feasibility of producing bio-ethanol from waste residues in Canada. Resources,
Conservation and Recycling, 50(3), 211–230.
Chandel, A. K., da Silva, S. S., Carvalho, W., & Singh, O. V. (2012). Sugarcane bagasse and
leaves: Foreseeable biomass of biofuel and bio-products. Journal of Chemical Technology &
Biotechnology, 87(1), 11–20.
Chaturvedi, V. K., Agarwal, S., Gupta, K. K., Ramteke, P. W., & Singh, M. P. (2018). Medicinal
mushroom: Boon for therapeutic applications. 3 Biotech, 8, 1–20.
Chen, H., & Chen, H. (2014). Biological fundamentals for the biotechnology of lignocellulose. In
Biotechnology of lignocellulose: Theory and practice (pp. 73–141). Springer.
Chi, Y., Hatakka, A., & Maijala, P. (2007). Can co-culturing of two white-rot fungi increase lignin
degradation and the production of lignin-degrading enzymes? International Biodeterioration
& Biodegradation, 59(1), 32–39.
Ćilerdžić, J., Stajić, M., & Vukojević, J. (2016). Degradation of wheat straw and oak sawdust by
Ganoderma applanatum. International Biodeterioration & Biodegradation, 114, 39–44.
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 283

Clutterbuck, A. J. (1990). The genetics of conidiophore pigmentation in Aspergillus nidulans.


Microbiology, 136(9), 1731–1738.
Elsacker, E., Vandelook, S., Van Wylick, A., Ruytinx, J., De Laet, L., & Peeters, E. (2020). A
comprehensive framework for the production of mycelium-based lignocellulosic composites.
Science of the Total Environment, 725, 138431.
Eriksson, K. E. L., Blanchette, R. A., Ander, P., Eriksson, K. E. L., Blanchette, R. A., & Ander,
P. (1990). Morphological aspects of wood degradation by fungi and bacteria. In Microbial and
enzymatic degradation of wood and wood components (pp. 1–87). Springer.
Falade, A. O., Nwodo, U. U., Iweriebor, B. C., Green, E., Mabinya, L. V., & Okoh, A. I. (2017).
Lignin peroxidase functionalities and prospective applications. Microbiology, 6, 1–14.
Fragoeiro, S., & Magan, N. (2005). Enzymatic activity, osmotic stress and degradation of pesti-
cide mixtures in soil extract liquid broth inoculated with Phanerochaete chrysosporium and
Trametes versicolor. Environmental Microbiology, 7(3), 348–355.
Galbe, M., & Zacchi, G. (2012). Pretreatment: The key to efficient utilization of lignocellulosic
materials. Biomass and Bioenergy, 46, 70–78.
García-Torreiro, M., López-Abelairas, M., Lu-Chau, T. A., & Lema, J. M. (2016). Fungal pretreat-
ment of agricultural residues for bioethanol production. Industrial Crops and Products, 89,
486–492.
Goodell, B., Qian, Y., & Jellison, J. (2008). Chapter 2: Fungal decay of wood: Soft rot—Brown
rot—White rot. In Development of commercial wood preservatives (pp. 9–31). American
Chemical Society.
Hamed, S. A. M. (2013). In-vitro studies on wood degradation in soil by soft-rot fungi: Aspergillus
niger and Penicillium chrysogenum. International Biodeterioration & Biodegradation,
78, 98–102.
Hamzeh, Y., Ashori, A., Khorasani, Z., Abdulkhani, A., & Abyaz, A. (2013). Pre-extraction of
hemicelluloses from bagasse fibers: Effects of dry-strength additives on paper properties.
Industrial Crops and Products, 43, 365–371.
Hasunuma, T., Okazaki, F., Okai, N., Hara, K. Y., Ishii, J., & Kondo, A. (2013). A review of
enzymes and microbes for lignocellulosic biorefinery and the possibility of their application to
consolidated bioprocessing technology. Bioresource Technology, 135, 513–522.
Hatakka, A., & Hammel, K. E. (2011). Fungal biodegradation of lignocelluloses. In Industrial
applications (pp. 319–340). Springer.
Hemmilä, V., Adamopoulos, S., Karlsson, O., & Kumar, A. (2017). Development of sustainable
bio-adhesives for engineered wood panels–A review. RSC Advances, 7(61), 38604–38630.
Huang, D. L., Zeng, G. M., Feng, C. L., Hu, S., Zhao, M. H., Lai, C., Zhang, Y., Jiang, X. Y., & Liu,
H. L. (2010). Mycelial growth and solid-state fermentation of lignocellulosic waste by white-­
rot fungus Phanerochaete chrysosporium under lead stress. Chemosphere, 81(9), 1091–1097.
Jones, M., Mautner, A., Luenco, S., Bismarck, A., & John, S. (2020). Engineered mycelium com-
posite construction materials from fungal biorefineries: A critical review. Materials & Design,
187, 108397.
Kamat, S., Khot, M., Zinjarde, S., RaviKumar, A., & Gade, W. N. (2013). Coupled production of
single cell oil as biodiesel feedstock, xylitol and xylanase from sugarcane bagasse in a biore-
finery concept using fungi from the tropical mangrove wetlands. Bioresource Technology, 135,
246–253.
Kameshwar, A. K. S., & Qin, W. (2016). Lignin degrading fungal enzymes. In Production of bio-
fuels and chemicals from lignin (pp. 81–130). Springer.
Kersten, P., & Cullen, D. (2007). Extracellular oxidative systems of the lignin-degrading
Basidiomycete Phanerochaete chrysosporium. Fungal Genetics and Biology, 44(2), 77–87.
Kirk, T. K. (1975). Effects of a brown-rot fungus, Lenzites trabea, on lignin in spruce wood.
Holzforschung, 29, 99.
Kuhad, R. C., Kuhar, S., Kapoor, M., Sharma, K. K., & Singh, A. (2007). Lignocellulolytic micro-
organisms, their enzymes and possible biotechnologies based on lignocellulolytic microorgan-
isms and their enzymes. In Lignocellulose biotechnology: Future prospects (pp. 3–22). IK
International Pvt Ltd.
284 S. Chepyala et al.

Kuhad, R. C., Gupta, R., & Singh, A. (2011). Microbial cellulases and their industrial applications.
Enzyme Research, 2011, 280696.
Kuila, A., & Sharma, V. (Eds.). (2017). Lignocellulosic biomass production and industrial applica-
tions (pp. 1–276). Wiley.
Kumar, R., Singh, S., & Singh, O. V. (2008). Bioconversion of lignocellulosic biomass: Biochemical
and molecular perspectives. Journal of Industrial Microbiology and Biotechnology, 35(5),
377–391.
Kumar, A., Gautam, A., & Dutt, D. (2020). Bio-pulping: An energy saving and environment-­
friendly approach. Physical Sciences Reviews, 5(10), 20190043.
Lombard, V., Golaconda Ramulu, H., Drula, E., Coutinho, P. M., & Henrissat, B. (2014). The
carbohydrate-active enzymes database (CAZy) in 2013. Nucleic Acids Research, 42(D1),
D490–D495.
Mantanis, G. I., Athanassiadou, E. T., Barbu, M. C., & Wijnendaele, K. (2018). Adhesive sys-
tems used in the European particleboard, MDF and OSB industries. Wood Material Science &
Engineering, 13(2), 104–116.
Marđetko, N., Trontel, A., Novak, M., Pavlečić, M., Ljubas, B. D., Grubišić, M., Tominac, V. P.,
Ludwig, R., & Šantek, B. (2021). Screening of lignocellulolytic enzyme activities in fungal
species and sequential solid-state and submerged cultivation for the production of enzyme
cocktails. Polymers, 13(21), 3736.
Marinovíc, M., Di Falco, M., Aguilar Pontes, M. V., Gorzsás, A., Tsang, A., de Vries, R. P.,
Mäkelä, M. R., & Hildén, K. (2022). Comparative analysis of enzyme production patterns of
lignocellulose degradation of two white rot fungi: Obba rivulosa and Gelatoporia subvermis-
pora. Biomolecules, 12(8), 1017.
Martínez, Á. T., Speranza, M., Ruiz-Dueñas, F. J., Ferreira, P., Camarero, S., Guillén, F., Martínez,
M. J., Gutiérrez Suárez, A., & Río Andrade, J. C. D. (2005). Biodegradation of lignocellulos-
ics: Microbial, chemical, and enzymatic aspects of the fungal attack of lignin. International
Microbiology, 8(3), 195–204.
Masarin, F., Pavan, P. C., Vicentim, M. P., Souza-Cruz, P. B., Loguercio-Leite, C., & Ferraz,
A. (2009). Laboratory and mill scale evaluation of biopulping of Eucalyptus grandis
Hill ex Maiden with Phanerochaete chrysosporium RP-78 under non-aseptic conditions.
Holzforschung, 63(3), 259–263.
Mayer, A. M., & Staples, R. C. (2002). Laccase: New functions for an old enzyme. Phytochemistry,
60(6), 551–565.
Misra, S., Raghuwanshi, S., & Saxena, R. K. (2013). Evaluation of corncob hemicellulosic hydro-
lysate for xylitol production by adapted strain of Candida tropicalis. Carbohydrate Polymers,
92(2), 1596–1601.
Motta, F. L., Andrade, C. C. P., & Santana, M. H. A. (2013). A review of xylanase production
by the fermentation of xylan: Classification, characterization and applications. In Sustainable
degradation of lignocellulosic biomass-techniques, applications and commercialization (Vol.
1, pp. 251–276). InTech.
Padma, N., & Sudha, H. (2013). Optimization of lignin peroxidase, manganese peroxidase, and
Lac production from Ganoderma lucidum under solid state fermentation of pineapple leaf.
BioResources, 8, 250–271.
Parani, K., & Eyini, M. (2010). Effect of co-fungal treatment on biodegradation of coffee pulp
waste in solid state fermentation. Asian Journal of Experimental Biological Sciences, 1(2),
352–359.
Pramanik, K., & Sahu, S. (2017). Biological treatment of lignocellulosic biomass to bioethanol.
Advances in Biotechnology and Microbiology, 5, 1–3.
Ragauskas, A. J., Williams, C. K., Davison, B. H., Britovsek, G., Cairney, J., Eckert, C. A.,
Frederick, W. J., Jr., Hallett, J. P., Leak, D. J., Liotta, C. L., & Mielenz, J. R. (2006). The path
forward for biofuels and biomaterials. Science, 311(5760), 484–489.
Reddy, G. V., Babu, P. R., Komaraiah, P., Roy, K. R. R. M., & Kothari, I. L. (2003). Utilization
of banana waste for the production of lignolytic and cellulolytic enzymes by solid substrate
Myco-degradation of Lignocellulosic Waste Biomass and Their Applications 285

fermentation using two Pleurotus species (P. ostreatus and P. sajor-caju). Process Biochemistry,
38(10), 1457–1462.
Renvall, P. (1995). Community structure and dynamics of wood-rotting Basidiomycetes on decom-
posing conifer trunks in northern Finland. Karstenia, 35, 1–51.
Robl, D., Delabona, P. D. S., Mergel, C. M., Rojas, J. D., Costa, P. D. S., Pimentel, I. C., Vicente,
V. A., da Cruz Pradella, J. G., & Padilla, G. (2013). The capability of endophytic fungi for pro-
duction of hemicellulases and related enzymes. BMC Biotechnology, 13(1), 1–12.
Rohowsky, B., Häßler, T., Gladis, A., Remmele, E., Schieder, D., & Faulstich, M. (2013). Feasibility
of simultaneous saccharification and juice co-fermentation on hydrothermal pretreated sweet
sorghum bagasse for ethanol production. Applied Energy, 102, 211–219.
Sakamoto, T., Hasunuma, T., Hori, Y., Yamada, R., & Kondo, A. (2012). Direct ethanol production
from hemicellulosic materials of rice straw by use of an engineered yeast strain codisplaying
three types of hemicellulolytic enzymes on the surface of xylose-utilizing Saccharomyces cere-
visiae cells. Journal of Biotechnology, 158(4), 203–210.
Sánchez, C. (2009). Lignocellulosic residues: Biodegradation and bioconversion by fungi.
Biotechnology Advances, 27(2), 185–194.
Saritha, M., Arora, A., & Lata. (2012). Biological pretreatment of lignocellulosic substrates for
enhanced delignification and enzymatic digestibility. Indian Journal of Microbiology, 52,
122–130.
Schwarze, F. W., Engels, J., & Mattheck, C. (2000). Fungal strategies of wood decay in trees.
Springer Science & Business Media.
Segato, F., Damásio, A. R. L., de Lucas, R. C., Squina, F. M., & Prade, R. A. (2014). Genome
analyses highlight the different biological roles of cellulases. Microbiology and Molecular
Biology Reviews, 78, 588–613.
Serrano-Ruiz, J. C., Luque, R., & Sepúlveda-Escribano, A. (2011). Transformations of biomass-­
derived platform molecules: From high added-value chemicals to fuels via aqueous-phase pro-
cessing. Chemical Society Reviews, 40(11), 5266–5281.
Shahsavarani, H., Hasegawa, D., Yokota, D., Sugiyama, M., Kaneko, Y., Boonchird, C., &
Harashima, S. (2013). Enhanced bio-ethanol production from cellulosic materials by semi-­
simultaneous saccharification and fermentation using high temperature resistant Saccharomyces
cerevisiae TJ14. Journal of Bioscience and Bioengineering, 115(1), 20–23.
Sindhu, R., Binod, P., & Pandey, A. (2016). Biological pretreatment of lignocellulosic biomass–An
overview. Bioresource Technology, 199, 76–82.
Singh, P., Sulaiman, O., Hashim, R., Peng, L. C, & Singh, R. P. (2013). Using biomass residues
from oil palm industry as a raw material for pulp and paper industry: potential benefits and
threat to the environment. Environment, development and sustainability, 15, pp.367–383.
Solt, P., Konnerth, J., Gindl-Altmutter, W., Kantner, W., Moser, J., Mitter, R., & van Herwijnen,
H. W. (2019). Technological performance of formaldehyde-free adhesive alternatives for par-
ticleboard industry. International Journal of Adhesion and Adhesives, 94, 99–131.
Songulashvili, G., Elisashvili, V., Wasser, S., Nevo, E., & Hadar, Y. (2006). Laccase and manga-
nese peroxidase activities of Phellinus robustus and Ganoderma adspersum grown on food
industry wastes in submerged fermentation. Biotechnology Letters, 28, 1425–1429.
Tani, S., Kawaguchi, T., & Kobayashi, T. (2014). Complex regulation of hydrolytic enzyme
genes for cellulosic biomass degradation in filamentous fungi. Applied Microbiology and
Biotechnology, 98, 4829–4837.
Tuomela, M., Vikman, M., Hatakka, A., & Itävaara, M. (2000). Biodegradation of lignin in a com-
post environment: A review. Bioresource Technology, 72(2), 169–183.
Van Dyk, J. S., & Pletschke, B. (2012). A review of lignocellulose bioconversion using enzymatic
hydrolysis and synergistic cooperation between enzymes—Factors affecting enzymes, conver-
sion and synergy. Biotechnology Advances, 30(6), 1458–1480.
Van Gool, M. P., Toth, K., Schols, H. A., Szakacs, G., & Gruppen, H. (2012). Performance of
hemicellulolytic enzymes in culture supernatants from a wide range of fungi on insoluble
wheat straw and corn fiber fractions. Bioresource Technology, 114, 523–528.
286 S. Chepyala et al.

Vantamuri, A. B., & Kaliwal, B. B. (2016). Production of laccase by newly isolated Marasmius sp.
BBKAV79 in solid state fermentation and its antiproliferative activity. International Journal of
Pharmaceutical Sciences and Research, 7(12), 4978.
Walter, M., & Boyd-Wilson, K. S. (2006). Development of a biotechnology tool using New Zealand
white-rot fungi to degrade pentachlorophenol in soil–A summary. In Modern multidisciplinary
applied microbiology: Exploiting microbes and their interactions (pp. 55–59). Wiley.
Wan, C., & Li, Y. (2012). Fungal pretreatment of lignocellulosic biomass. Biotechnology Advances,
30(6), 1447–1457.
Wang, J., Li, L., Xu, H., Zhang, Y., Liu, Y., Zhang, F., Shen, G., Yan, L., Wang, W., Tang, H., & Qiu,
H. (2022). Construction of a fungal consortium for effective degradation of rice straw lignin
and potential application in bio-pulping. Bioresource Technology, 344, 126168.
Wen, Z., Liao, W., & Chen, S. (2004). Hydrolysis of animal manure lignocellulosics for reducing
sugar production. Bioresource Technology, 91(1), 31–39.
Wong, D. W. (2009). Structure and action mechanism of ligninolytic enzymes. Applied
Biochemistry and Biotechnology, 157, 174–209.
Yamada, R., Taniguchi, N., Tanaka, T., Ogino, C., Fukuda, H., & Kondo, A. (2010). Cocktail
δ-integration: A novel method to construct cellulolytic enzyme expression ratio-optimized
yeast strains. Microbial Cell Factories, 9, 1–8.
Yu, H., Guo, G., Zhang, X., Yan, K., & Xu, C. (2009a). The effect of biological pretreatment with
the selective white-rot fungus Echinodontium taxodii on enzymatic hydrolysis of softwoods
and hardwoods. Bioresource Technology, 100(21), 5170–5175.
Yu, M., Zeng, G., Chen, Y., Yu, H., Huang, D., & Tang, L. (2009b). Influence of Phanerochaete
chrysosporium on microbial communities and lignocellulose degradation during solid-state
fermentation of rice straw. Process Biochemistry, 44(1), 17–22.
Zhao, X., Huang, X., Yao, J., Zhou, Y., & Jia, R. (2015). Fungal growth and manganese peroxidase
production in a deep tray solid-state bioreactor, and in vitro decolorization of poly R-478 by
MnP. Journal of Microbiology and Biotechnology, 25(6), 803–813.
Zhou, C. H. C., Beltramini, J. N., Fan, Y. X., & Lu, G. M. (2008). Chemoselective catalytic conver-
sion of glycerol as a biorenewable source to valuable commodity chemicals. Chemical Society
Reviews, 37(3), 527–549.
Value-Added Product Development
Utilising the Food Wastes

Anduri Sravani, C. R. Patil, and Shivani Sharma

1 Introduction

Food is one among the primary source of survival; problems such as “food loss” and
“food waste” should be seriously considered and addressed in order to reduce their
prevalence and the potential effects they may have not only on humans but also on
the ecosystem and the overall balance of the environment. The counter “inedible
parts” that accompany “food” are common and may or may not be suitable for
usage (Lin et al., 2022). The global population is expanding leading to an increase
in both the amount of food requirement and the amount of agricultural and food
waste produced. The Food and Agriculture Organisation of the United Nations
(FAO) defines food waste as a reduction in the quality or quantity of food brought
on by the choices and deeds of retailers, food service providers and customers.
According to the FAO, 14% of the food produced worldwide or $400 billion worth
food is wasted yearly after harvest but before it reaches shops. Additionally, 17% of
food waste is generated from retailers and from consumers, especially homes as per
the reports of Food Waste Index from the United Nations Environment Programme.
Food wastes are of three kinds such as solid food waste, semi-solid food waste
and liquid food waste. Solid wastes mostly consist of the food products themselves,
their components or other products connected to a specific food product. These
products are typically high in starch, lignin, cellulose and monosaccharides (Yahia

A. Sravani (*)
Department of Agricultural Microbiology, College of Agriculture, University of Agricultural
Sciences, Dharwad, India
C. R. Patil
Department of Agricultural Microbiology, Institute of Organic Farming, University of
Agricultural Sciences, Dharwad, India
S. Sharma
Department of Microbiology, Punjab Agricultural University, Ludhiana, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 287
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_16
288 A. Sravani et al.

& Carrillo-Lopez, 2018). According to data from Seberini (2020) and observations
from the FAO, approximately 46% of the fruit waste is generated including vegeta-
bles, roots and tubers, oranges and pineapples. On the contrary, 35% of seafood
including fish are wasted and 30% of cereals are wasted. According to data that the
FAO collected and analysed, this amount represents the food that is produced
approximately on 28% of the agricultural area of the world (1.4 billion hectares of
land globally). Hence, it is better to prevent food waste; nonetheless, a large amount
of food waste is unavoidable due to the growing global population.
The prevention of food waste generation and its recycling by valorisation are the
solutions to reduce the food wastes. Utilising a sustainable process called valorisa-
tion, food waste can be turned into goods with value-added products based on its
bioactive ingredients. These bioactive ingredients, which include polyphenols, vita-
mins, mineral and prebiotics not only enhance the value of the products but also
provide health benefits (Vilas-Boas et al., 2021). It is feasible to develop premium
and useful food ingredients, cosmetics and nutritional supplements by utilising food
wastes for extracting these substances. In addition to reducing food waste, this strat-
egy intends to provide new revenue streams and promote the development of a cir-
cular economy. Future food waste valorisation must take into account this waste’s
availability over time, its techno-economic potential and other factors in order to be
both environmentally and economically sustainable.

2 Food Losses and Food Wastes

Food loss, according to the FAO, is a reduction in the weight or nutritional content
of food that was basically used for human consumption. Food waste is defined as
food that is wasted even after it is kept over its expiry date or allowed to deteriorate
even though it is suitable for consumption by humans (FAO, 2013). Problems with
the institutional and policy framework that supports the food production and supply
chain are mostly to blame for food loss. In addition to this, administrative and tech-
nological limitations such as lack of appropriate storage facilities, a working cold
chain, acceptable food handling procedures, infrastructure, packaging or efficient
marketing strategies are also the reasons for food loss. It also includes food lost as
a result of pests and diseases. Food waste is defined as food that is removed from the
food supply chain despite still being fit for human consumption either deliberately
or after the food has deteriorated or expired as a result of bad inventory management
or carelessness. Food loss occurs majorly during the early stages of the food supply
chain system (during production stage, post-harvest stage and also during process-
ing stage). While, food waste occurs frequently at the final stages of food supply
chain system (retailers and consumers stage). According to Brian et al. (2013), the
terms “food loss” and “food waste” refer to different aspects of the same problem,
they are frequently used interchangeably. Food loss including post-harvest losses
and production losses refers to the unintentional degradation of food quality and
Value-Added Product Development Utilising the Food Wastes 289

quantity prior to consumption. Put simply, food waste and food loss are harmful for
the state of food security and have an impact on the economy by assuming pro-
longed pricing (Ishangulyyev et al., 2019).
Millions of people around the world are malnourished and the demand for food
is increasing. Globally, 1/3rd of food production which accounts for approximately
1.3 billion tonnes is lost or squandered in the food supply chain. With 1.3 billion
people, India generates 0.5 kg of organic garbage per person per day (Paulraj et al.,
2019). In India, significant amounts of food are wasted at hotels, shops, apartments,
restaurants, cafeterias on aeroplanes and the food processing industry. As per the
reports of Food Waste Index Report 2021 (United Nations Environment
Programme’s), in India, 90 kg of food waste per capita per year is in the high income
category, while it was 68 and 63 in the middle and low income sectors, respectively.
Food loss and waste can happen at any stage of the food supply chain including pre-­
harvest, harvest, breeding, handling, storage and transport afterward; processing
such as canning, packaging and transformation; distribution such as retail, transport
and consumption such as preparation and serving (Thyberg et al., 2015). Food waste
is produced in significant amounts at the retailer level. Retailers often have strict
standards for appearance when it comes to food. Because of this, fruits and vegeta-
bles that seem damaged are frequently not put on the display. The fishing industry
wastes a lot of food; between 40% and 60% of fish are thrown away in Europe as
they are of wrong kind or size (Stuart, 2009). The dairy industry is one of the kind
of the food industry that is present all over the world. In addition to producing a
sizable volume of solid and liquid waste, it also produces a wide variety of goods
such as milk, milk powder, butter and cheese (Jaganmai & Jinka, 2017). Food
wastes should be converted in to value-added products by valorisation. Food waste
valorisation must take into account the timely availability of this waste around the
year, its technological, economic aspects based on its life cycle in order to be both
environmentally and economically sustainable.

3 International and National Scenario of Food Wastes

For the creation of well-thought-out and successful policies and programmes, which
can be used to notice the modifications in residual flows after food loss and waste
prevention and recovery policies are executed, it is crucial to quantify the amount of
food loss and wastes (Thyberg et al., 2015). Understanding food loss and wastes can
inspire people to alter their attitudes and behaviours. According to FSSAI’s most
recent figures (2022), India is the second largest food producer globally, accounting
for close to 10.1% of total food produced globally. Food and Agricultural
Organisation estimates that 25% of the world’s population suffers from hunger, and
statistical studies have shown that as of 2021, India’s food waste accounted for close
to 40% of its overall food production on weight basis, which includes both grains
290 A. Sravani et al.

and other foods. This includes domestic waste from each person throwing out an
average of 50 kg of food annually (Roe et al., 2021).
Globally, food loss and wastes account for 1.6 billion tons annually (614 kcal/
person each day). Food loss and wastes in Australia account for 4.06 million tons
annually, China accounts for 90 million tons (51% of total municipal solid waste),
Denmark accounts for 700,000 tons annually and Finland accounts for 23 kg per-
son−1 year−1. Italy accounts for 17.7 tons which accounts for 3.25% of its overall
production, Japan accounts for 37.86 million tons during 2011, New Zealand
accounts for 148 kg household−1 year−1. While, Nordic countries account for
40,000–83,000 tons yearly, Singapore accounts for 809,800 tons during 2017. South
Africa accounts for 177 kg person−1 yearly and Switzerland accounts for 48% of
total calories. Moreover, the United Kingdom accounts for 4.2 million tons yearly
and the United States accounts for 35.5 million tons annually (Ishangulyyev et al.,
2019). The large quantity of food waste is categorised into various food categories
based on the influence of environment on the kind of food. The high food waste
generation commodities to low food waste generation commodities are given in the
order as 44% (fruits and vegetables), 20% (roots and tubers), 19% (cereals), 8%
(milk), 4% (meat), 3% (oilseeds and pulses) and 2% (seafood) (Lipinski et al., 2013).

4 Nutritional Content of Food Waste

Food wastes are nothing but the highly organic residues formed from food handling
to processing either in solid or liquid form. These wastes are rich in proteins, carbo-
hydrates, fats, fibres, micronutrients and bioactive compounds and hence can form
base for the production of high-value products. The nutrients in food wastes are
distributed unevenly. In 80% of the foods dumped in roads in 2016 and also in 2018,
carbohydrates were the predominant nutrient and generalised linear mixed model
(GLMM) results showed that this dominance relative to the other nutrients was
significant in both 2016 and also in 2018. During both years, carbohydrates were the
most prevalent nutrient in foods sold in parks. Protein was the second most frequent
nutrient to predominate in food products throughout both years and site types such
as parks and median sites. Generalised linear mixed models (GLMMs) indicated
that they were only marginally less prevalent in both site categories in 2016.
However, in medians, it was just somewhat less dominating in 2018. Only in 2016
medians and 2018 parks were lipid-dominated foods discovered and even then, they
only made up about 5% of the foods actually observed (Carpenter & Savage, 2021).
Carbohydrate content in surveyed foods was considerably higher than fat and
protein content in medians and parks in both 2016 and 2018. The relative content of
carbohydrates was much higher than the other nutrients for both years and both site
types. In 2016, food items identified in medians had an average carbohydrate con-
tent of 54 g/100 g of food. Food items had an average of 32 g of carbohydrates per
100 g in parks and medians in 2018. Comparatively, it was discovered that the levels
Value-Added Product Development Utilising the Food Wastes 291

of proteins and lipids were much lower in both the years and the site types. In 2016,
medians had an average protein content of 5 g, parks had an average protein content
of 4 g, and both site types had an average lipid content of 4 g. In 2018, medians had
an average of 5.5 g of protein and parks had an average of 10 g; medians had an
average of 4 g of lipids and parks had an average of 7 g (Carpenter & Savage, 2021).

5 Conversion Methods for Food Waste into By-Products

Food and Agriculture Organization of the United Nations (FAO (2013) published
facts regarding food wastes, such as consumers in wealthy nations annually squan-
der nearly as much food, that is, 222 million tonnes as total net food production of
Africa is 230 million tonnes. Memon (2010) proposed the 3R’s concept, that is,
Reduce, Reuse, and Recycle to address the global problem of food waste.
Development of infrastructure, transportation, food processing and packaging
industries can all profit from value-added product generation. This has not only
major financial benefits but also helps to reduce trash generation. In countries in
Africa, Asia and America, tropical fruits, dairy products and fish are significant
commercial food and crop industries that contribute significantly to the socioeco-
nomic growth of rural and urban populations. The following are the four sugges-
tions made by Riemer and Kristoffersen (1999) for ideal waste minimisation in the
sector as follows:
1. Waste reduction through the use of more productive production techniques.
2. Internal recycling of waste.
3. Improvement of waste quality focused on sources.
4. Product reuse. In the food business, a variety of by-products are also mostly
dumped without any further processing, despite the fact that they can be used to
create worthwhile goods. The approaches that have historically been preferred to
transform by-products into value-added goods are examined in this article.

5.1 Thermal Conversion

The thermal conversion of food waste or by-products, particularly solid waste, relies
on the creation of fuel and chemicals, which are typically utilised in power steam
turbines for the generation of energy or to heat process streams in industrial settings.
In contrast to pyrolysis, hydrothermal carbonisation uses autogenous pressures and
relatively low temperatures (180–350 °C) to transform food wastes into a lucrative,
energy-rich resource (Pham et al., 2015). For the purpose of removing textile colours
from contaminated water, Parshetti et al. (2014) prepared hydrochar from urban
food wastes using the hydrothermal carbonisation process. Pyrolysis and
292 A. Sravani et al.

hydrothermal carbonisation are the two fundamental thermal conversion processes.


Food waste burns at less than 450 °C and turns gaseous at more than 800 °C.

5.2 Chemical Conversion

For the most part, hydrolysis and oxidation reactions are used in the industries of
food processing as a means of chemically converting food wastes and by-products.
Green extraction techniques can also be used to extract valuable components from
food wastes and by-products, with water being favoured over organic solvent extrac-
tion as the extraction medium. For the recovery of antioxidant bioactive compounds
from winery wastes and by-products, the alternative extraction techniques such as
pulsed electric fields, high-voltage electrical discharges, pulsed ohmic heating,
ultrasounds, microwave- assisted extractions, sub and supercritical fluid extractions
as well as pressurised liquid extraction were also studied in depth (Barba et al.,
2016). Water was also employed to extract melon seed milk from melon seeds by
Zungur Bastıoglu et al. (2016) which are typically referred to as waste. With the
help of ethanol, Amado et al. (2014) attempted to extract antioxidants from leftover
potato peel. They explored how the temperature, duration and ethanol concentration
of the extraction procedure affected the extraction of antioxidants from potato peel.
Goula and Lazarides (2015) described how they valorised pomegranate and olive
mill refuse. They suggested several strategies for completely utilising pomegranate
seeds and peels to extract oil and phenolics from them and also for converting milled
olive waste into spread paste and olive powder mostly utilised in formulations of
food and polyphenols that are encapsulated.

5.3 Biological Conversion

Worldwide interest is currently growing in the bio treatment of waste food and their
by-products for extraction of energy and bioactive substances. Food waste contains
a lot of moisture and organic material, making it the ideal substrate for anaerobic
digestion. Hydrolysis, acidogenesis, acetogenesis and methanogenesis are the four
stages of anaerobic digestion. According to Jiang et al. (2013), PH, temperature and
the rate of organic loading all have an impact on the acidogenesis of food waste.
According to Chandrasekaran (2012), anaerobic digestion produces biogas, which
is mostly composed of CH4 and CO2 and can be utilised as a fuel similar to natural
gas. Composting of food scraps and by-products is another biological conversion
technique in addition to anaerobic digestion. The process of composting is biologi-
cal. Different techniques used for extraction of bioactive compounds from food
wastes are represented in Table 1.
Value-Added Product Development Utilising the Food Wastes 293

Table 1 Various methods used for extraction of bioactive compounds from food wastes
Extraction technique Bioactive compounds extracted
Food Solvent extraction Anthocyanin, lycopenes, polyphenols and Kumar et al.
wastes carotenoids (2017)
Microwave-assisted Catechins, mangiferin, phenols, polyphenols,
extraction saponins
Subcritical water Caffeic acid, chlorogenic acid, coumaric acid,
extraction ferulic acid, gallic acid, mangiferin
Supercritical fluid Flavonoids, polyphenols, tocopherols,
extraction epicatechin, catechin, procyanidines
Enzyme-assisted Alkaloids, coumarins, flavanoids, lectins,
extraction terpenoids
Ultrasound Anthocyanin, saponins, catechin, flavonoids,
extraction polyphenols, β-glucan

6 Production of Value-Added Products from Food Waste

Valorisation is the process of converting wastes into more value-added products


such as enzymes, bioactive compounds and fuels (Arancon et al., 2013). Food waste
valorisation uses first-generation and second-generation techniques and adoption
has been gradual. Utilising waste materials completely for the production of animal
feed, energy and compost for specific consumer applications is the goal of the first-­
generation valorisation process. Different types of fractionated usage of materials
are included in the second-generation valorisation process. To create various types
of industrial products such as chemicals, functional foods and commodity goods,
they rely on the integration of customised recovery and enhanced conversion tech-
niques for certain components. These cutting-edge conversion processes for partic-
ular food waste components are referred to as advanced second-generation
processing. The low disposal costs of the traditional approaches account for the
gradual shift towards valorisation (Xiong et al., 2019). Among the widespread first-­
generation valorisation techniques utilised worldwide are composting, methane col-
lection and energy production. These first-generation valorisation techniques work
well, but they are significantly more expensive than sophisticated second-generation
food waste processing (Arancon et al., 2013).
Bioactive substances are molecules with the potential to promote health that are
found in food in trace amounts. Bioactive substances more specifically have a
favourable impact on the human body or particular tissues or cells. It has been estab-
lished that numerous substances found in goods with both plant- and animal-derived
origins have favourable impacts on human health. Long-chain polyunsaturated fatty
acids (PUFA), vitamins, carotenoids, peptides and polyphenols are examples of bio-
active substances. Food waste items made from meat or fish may also contain bioac-
tive compounds because meat and other meat products contain a large amount of
bioactive compounds. For the production of nutraceuticals, functional foods and
food additives, bioactive substances make up an ideal pool of molecules (Joana Gil-­
Chávez et al., 2013). The most basic type of functional foods are fruits and
294 A. Sravani et al.

vegetables because they are abundant in a variety of bioactive compounds. According


to Day et al. (2009), fruits high in polyphenols and carotenoids have antioxidant
properties that lower the chance of getting certain cancers. Vegetable trimmings,
peelings, stalks, seeds, shells, bran, leftover juice, oil, starch and sugar residues are
all considered to be vegetable waste. Garbage from the dairy and seafood industries
is included in the category of animal-derived garbage. The recovered biomolecules
and by-products can be utilised in the food industry, pharmaceutical and medical
preparations and the production of functional foods (Baiano, 2014). Significant
amounts of antioxidant activity can be found in bioactive phytochemicals such ste-
rols, tocopherols, carotenes, terpenes and polyphenols that are isolated from tomato
by-products. As a result, these value-adding components that were extracted from
such waste can be employed to create functional foods as natural antioxidants or as
additives to increase the shelf life of other foods (Kalogeropoulos et al., 2012).
Literature survey stated that bioactive compounds are majorly extracted by using
techniques such as subcritical water extraction, supercritical fluid extraction, sol-
vent extraction, utilising enzymes, microwaves and ultrasounds.
Pharmaceutical compounds such as pills, capsules, tablets, powder and vials are
the most common forms in which nutraceuticals are used (Espín et al., 2007). The
1,3,6,7-tetrahydroxyxanthone-C2-d-glucoside (Mangiferin), a naturally occurring
bioactive xanthonoid found in several plant species such as the mango tree
(Mangifera indica L), has drawn the attention of several research teams globally for
its potential to treat cancer according to Núñez Selles et al. (2016). Mangiferin has
demonstrated potential benefits in leukaemia, brain, lung, cervical, breast and pros-
tate cancers in addition to its antioxidant and anti-inflammatory effects when given
alone or in conjunction with recognised anticancer drugs. After being extracted,
cholesterol from by-products of the meat industry including brains, nervous sys-
tems and spinal cords is utilised to create vitamin D3 (Ejike & Emmanuel, 2009).
Volarisation of food waste using solid-state fermentation results in the produc-
tion of several enzymes such as cellulases, amylases, lipases, proteases, pectinases,
laccases and xylanases. Solid-state fermentation provides several advantages such
as low input cost, higher yields, simple equipment, less waste generation, culture
media obtained from organic, solid agricultural waste. Different microbial strains
such as Geobacillus sp., Bacillus amyloliquefaciens produced amylases with high
activity under optimum fermentation conditions utilising agricultural residues (rice
husks, corn cobs, coffee waste and tomato pomace), food wastes and kitchen waste
(potato peels and watermelon rinds). Iqbalsyah et al. (2019) studied the amylase
production with rice husk as a substrate by solid-state fermentation using Geobacillus
sp. with highest amylase activity of 1.85 U g−1 at 48 h. Few bacterial strains of the
Bacillus genus and several fungal strains such as Rhizopus, Aspergillus and
Penicillium have been found to be the highest producers of most active proteases.
Camargo et al. (2022) utilised processed wastes of oranges and grapes for estimat-
ing their bromatological properties and also for the production of proteases by using
Aspergillus niger CBMAI 2084. Lipases catalyse the synthesis of new products
through alcoholysis, acidolysis, esterification and transesterification methods
(Uçkun Kiran et al., 2014). Lipases are produced from lignocellulose-rich
Value-Added Product Development Utilising the Food Wastes 295

agricultural waste using several microorganisms such as Aspergillus niger and


Yarrowia lipolytica. A recent study investigated the production of lipase through
Penicillium roqueforti growth via SSF using cocoa bran residues (Araujo et al., 2022).
Only a few number of bacteria have been used for the commercial manufacture
of PHA, despite the fact that 250 different species of natural PHA producers have
been found. It has been discovered that these bacteria, which also include
Cupriavidus necator, Alcaligenes latus, Pseudomonas oleovorans and Bacillus
megaterium can convert various forms of carbon sources into poly hydroxyl alkano-
ates. One of the most often used microbial strains for the production of PHA is
C. necator, specifically (Reddy et al., 2003). According to the most recent studies,
the bacterium Halomonas hydrothermalis, H. campaniensis LS21 can grow in both
synthetic sea water and food wastes such as mixed substrates made of cellulose,
proteins, lipids, fatty acids, starch and produced approximately 70% polyhydroxy-
butyrate with a pH of 10 at 37 °C. Animal by-products such as milk, cheese fats and
intestinal waste are used as a substrates for the manufacture of PHA by using
microbes such as Alcaligenes latus and Bacillus megaterium. In two distinct fer-
mentation temperature setups, the experimental results revealed a maximum spe-
cific growth rate of 0.10 g L−1 h−1 (Muhr et al., 2013). The bioplastic’s volumetric
productivity arrived at 0.036 and 0.050 g.
Currently, edible food wastes are used to make a major portion of biofuels such
as biodiesel and bioethanol. Biodiesel is made from a variety of edible plant oils
including canola, rapeseed and soybean oils. As opposed to this, ethanol may be
made from a wide range of feedstocks including sugar cane, bagasse, sugar beet,
grain, switchgrass, barley, potatoes, molasses, corn, stover, wheat and many more
sources rich in carbohydrates (Pimentel & Patzek, 2005; Karmee & Chadha, 2005).
Transesterification is the chemical process used to make biodiesel. Tri, di and mono-
glycerides react with methanol in the presence of a catalyst during transesterifica-
tion to create biodiesel. Contrarily, the method for making bioethanol includes
pretreatment, enzymatic hydrolysis, fermentation and distillation processes.
Technologies are currently available for the industrial-scale generation of biodiesel
and bioethanol. In this regard, Saccharomyces cerevisiae H058 was used in a pilot-­
scale synthesis of ethanol from food waste, which has already been documented
(Yan et al., 2013). Potato peel waste was used as a substrate for bioethanol produc-
tion using biocatalytic methods (Yan et al., 2011). Kim et al. (2008) extracted the oil
from leftover noodles using nonpolar hexane as a solvent. From 100 g of leftover
noodles, they extracted about 83 g of refined starch and 5 g of oil. In order to pro-
duce bioethanol, the resulting oil-free starch residue underwent simultaneous sac-
charification and fermentation processes. Methane is created through anaerobic
digestion by biodegrading and reducing organic waste. According to Park et al.
(2018), using a microbial electrolysis cell (MEC) in addition to an anaerobic diges-
tion reactor increases the rate of methane production by 1.7 times when compared
to using an anaerobic digestion reactor alone. The fermentation of food waste
through biological process to produce hydrogen gas is called biohydrogen. Because
carbohydrate-rich food waste has 20 times greater potential than food waste rich in
protein or fat, carbohydrate-rich FW is chosen for the production of biohydrogen gas.
296 A. Sravani et al.

According to Gaur et al. (2019), biosurfactants are surface-active chemicals with


biological origins that are primarily produced by a variety of microorganisms
including bacteria, fungi and yeast. As a fermentation substrate for the synthesis of
biosurfactants, Pseudomonas aeruginosa isolated from kitchen waste oil preferred
to use kitchen waste oil over glucose, glycerol, molasses and rapeseed oil (Chen
et al., 2018). Production of paneer results in significant additional waste for the
Indian dairy industry. Two million tonnes of waste whey were produced each year
from the manufacture of 0.15 million tonnes of paneer. This waste whey is fre-
quently dumped into the environment without any pretreatment which pollutes the
soil and water (Parashar et al., 2016). Paneer whey was used by the hydrocarbon-­
contaminated soil strain Pseudomonas aeruginosa SR17 to produce 2.7 g/L of bio-
surfactant. A Pseudozyma species strain produced a biosurfactant that had potential
for use in laundry detergent (Sajna et al., 2013). According to Ramírez et al. (2015),
Pseudomonas aeruginosa and Bacillus subtilis used olive oil mill waste to create
biosurfactant.
In microbial fuel cell, microbes oxidise organic matter by sending electrons to
the anode, and at the cathode, oxidised chemicals or oxygen are reduced by microbes
or by an abiotic mechanism. The hydrolysis of this organic fraction is the rate-­
limiting stage in the production of electricity and the organic matter-rich food waste
serves as an energy source for electricigens in microbial fuel cell (Li et al., 2016).
According to reports, pretreatment of food waste using microwave and sonication
improves substrate hydrolysis which in turn boosts electricity production efficiency
(Yusoff et al., 2013). Maximum power of 15.14 W/m3 and open circuit voltage of
1.12 V were obtained by microbial fuel cell containing food waste leachate at a
concentration of 5000 mg/L. With rising substrate concentration of 20,000 mg/L,
power output decreased; this decrease was due to anode chamber microbial inhibi-
tion (Rikame et al., 2012).
As a substrate, agricultural waste has been used in the production of mushrooms.
It reduces the environmental burden, improves agricultural productivity and modi-
fies the bacterial ecology in the soil all at once. Due to their nutritional and carbon
content as well as their abundance in macronutrients (N, K, P, Ca and Mg) and
micronutrients (Fe, Cu, Zn and Al), residues from food waste biogas generation can
be employed as organic fertilisers or soil conditioners (Zhu et al., 2015). Several
methods, including anaerobic digestion, aerobic composting with microbes, chemi-
cal hydrolysis method (processing food waste by alkaline or acid hydrolysis at
600–1000 °C) and in situ degradation of natural organic matter can be used to create
organic fertilisers from food waste. It creates soluble biowaste compost (SBC),
digestate/soil conditioner, compost, deteriorated crop, minerals and liquid organic
fertilisers (Du et al., 2018).
Agriculture waste can be used as a natural, affordable replacement for many
high-value items. By utilising peels, seeds, oil cakes, field residues and bran,
microbes can easily feed upon agricultural waste to produce a variety of high end
products such as pigments, phytochemicals, antibiotics and different enzymes
(endoglucanase, glucosidase, amylase and glucoamylase) (Bogar et al., 2003).
Exopolysaccharide known as xanthan is used as a food additive in the food industry
Value-Added Product Development Utilising the Food Wastes 297

Fig. 1 Value-added products developed from various kinds of food wastes

and is thought to be created by Xanthomonas species using agricultural waste as a


substrate. Food additives are crucial for improving the technological capabilities of
food. Edible mushrooms such as Lentinula edodes and Pleurotus sp., can be grown
economically from lignocellulosic agricultural waste (wheat paddy, rice paddy,
banana leaves and cotton stalks) (Philippoussis, 2009). According to Sadh et al.
(2018), agroindustrial waste is a good carrier for solid-state fermentation and
enzyme immobilisation.
Fruit and vegetable wastes are processed by microorganisms into value-added
products such as fermented beverages, namely, fenny, vinegar, single-cell proteins
(Saccharomyces sp., Candida utilis, Endomycopsis fibuligera and Pichia burtonii),
polysaccharides, dietary fibre, polyphenols, bio-pigments (carotenoids), fragrances,
flavours (vanillin) and essential oils. Fruit and vegetable wastes undergo acidogenic
fermentation which results in the production of lactic acid (Wu et al., 2015). In
contrast, solid-state fermentation of wastes using unrefined enzyme combinations
produces succinic acid (Dessie et al., 2018) (Fig. 1).

7 Conclusion

Food is an essential good that accounts for a sizable portion of the organic waste
produced worldwide. When food waste is not properly managed, hazardous sub-
stances including greenhouse gases, nitrates and ammonia are released into the
environment. Currently, the processing of food waste by various treatments and
techniques provides both environmental protection and a sustainable method for the
creation of billion-dollar profits by transforming food waste into complex bioactive
compounds, nutraceuticals, bioplastics, biofertilisers, biofuels, electrical energy,
mushrooms, single-cell proteins, fermented beverages, pigments, fragrances,
298 A. Sravani et al.

flavours and essential oils. As food waste contains several important nutrients, their
valorisation helps in food waste management by value addition and generating
revenue.

References

Amado, I. R., Franco, D., Sánchez, M., Zapata, C., & Vázquez, J. A. (2014). Optimisation of anti-
oxidant extraction from Solanum tuberosum potato peel waste by surface response methodol-
ogy. Food Chemistry, 165, 290–299.
Arancon, R. A. D., Lin, C. S. K., Chan, K. M., Kwan, T. H., & Luque, R. (2013). Advances on waste
valorization: New horizons for a more sustainable society. Energy Science & Engineering,
1(2), 53–71.
Araujo, S. C., Ramos, M. R. M. F., do Espírito Santo, E. L., de Menezes, L. H. S., de Carvalho,
M. S., Tavares, I. M. D. C., Franco, M., & de Oliveira, J. R. (2022). Optimization of lipase
production by Penicillium roqueforti ATCC 10110 through solid-state fermentation using agro-­
industrial residue based on a univariate analysis. Preparative Biochemistry & Biotechnology,
52(3), 325–330.
Baiano, A. (2014). Recovery of biomolecules from food wastes—A review. Molecules, 19(9),
14821–14842.
Barba, F. J., Zhu, Z., Koubaa, M., Sant’Ana, A. S., & Orlien, V. (2016). Green alternative methods
for the extraction of antioxidant bioactive compounds from winery wastes and by-products: A
review. Trends in Food Science & Technology, 49, 96–109.
Bogar, B., Szakacs, G., Linden, J. C., Pandey, A., & Tengerdy, R. P. (2003). Optimization of
phytase production by solid substrate fermentation. Journal of Industrial Microbiology and
Biotechnology, 30(3), 183–189.
Brian, L., Craig, H., James, L., Lisa, K., Richard, W., & Tim, S., (2013). Reducing food loss
and waste e installment 2 of “Creating a Sustainable Food Future”. World Resource Institute.
Washington, DC.
Camargo, D. A., Pereira, M. S., dos Santos, A. G., & Fleuri, L. F. (2022). Isolated and fermented
orange and grape wastes: Bromatological characterization and phytase, lipase and protease
source. Innovative Food Science & Emerging Technologies, 77, 102978.
Carpenter, M., & Savage, A. M. (2021). Nutrient availability in urban food waste: Carbohydrate
bias in the Philadelphia–Camden urban matrix. Journal of Urban Ecology, 7(1), juab012.
Chandrasekaran, M. (Ed.). (2012). Valorization of food processing by-products. CRC Press.
Chen, C., Sun, N., Li, D., Long, S., Tang, X., Xiao, G., & Wang, L. (2018). Optimization and
characterization of biosurfactant production from kitchen waste oil using Pseudomonas aeru-
ginosa. Environmental Science and Pollution Research, 25, 14934–14943.
Day, L., Seymour, R. B., Pitts, K. F., Konczak, I., & Lundin, L. (2009). Incorporation of functional
ingredients into foods. Trends in Food Science and Technology, 20, 388–395.
Dessie, W., Zhang, W., Xin, F., Dong, W., Zhang, M., Ma, J., & Jiang, M. (2018). Succinic acid
production from fruit and vegetable wastes hydrolyzed by on-site enzyme mixtures through
solid state fermentation. Bioresource Technology, 247, 1177–1180.
Du, C., Abdullah, J. J., Greetham, D., Fu, D., Yu, M., Ren, L., Li, S., & Lu, D. (2018). Valorization
of food waste into biofertiliser and its field application. Journal of Cleaner Production, 187,
273–284.
Ejike, C. E. C. C., & Emmanuel, T. N. (2009). Cholesterol concentration in different parts of
bovine meat sold in Nsukka, Nigeria: Implications for cardiovascular disease risk. African
Journal of Biochemistry Research, 3(4), 095–097.
Espín, J. C., García-Conesa, M. T., & Tomás-Barberán, F. A. (2007). Nutraceuticals: Facts and
fiction. Phytochemistry, 68(22–24), 2986–3008.
Value-Added Product Development Utilising the Food Wastes 299

FAO. (2013). Food wastage footprint: Impacts on natural resources. FAO.


Food and Agriculture Organization of the United Nations (FAO). (2013). FAO statistics. Available
online at: http://www.fao.org. Accessed July 13, 2017.
Food Safety and Standards Authority of India (FSSAI). (2022). An important outlook on adopting
it as a business. International Journal of Law Management & Humanities, 5(2), 246.
Gaur, V. K., Regar, R. K., Dhiman, N., Gautam, K., Srivastava, J. K., Patnaik, S., Kamthan, M.,
& Manickam, N. (2019). Biosynthesis and characterization of sophorolipid biosurfactant by
Candida spp.: Application as food emulsifier and antibacterial agent. Bioresource Technology,
285, 121314.
Goula, A. M., & Lazarides, H. N. (2015). Integrated processes can turn industrial food waste into
valuable food by-products and/or ingredients: The cases of olive mill and pomegranate wastes.
Journal of Food Engineering, 167, 45–50.
Iqbalsyah, T. M., Amna, U., Utami, R. S., & Oesman, F. (2019). Concomitant cellulase and amy-
lase production by a thermophilic bacterial isolate in a solid-state fermentation using rice
husks. Agriculture and Natural Resources, 53(4), 327–333.
Ishangulyyev, R., Kim, S., & Lee, S. H. (2019). Understanding food loss and waste—Why are we
losing and wasting food? Food, 8(8), 297.
Jaganmai, G., & Jinka, R. (2017). Production of lipases from dairy industry wastes and its applica-
tions. International Journal of Current Microbiology and Applied Sciences, 5, 67–73.
Jiang, J., Zhang, Y., Li, K., Wang, Q., Gong, C., & Li, M. (2013). Volatile fatty acids production
from food waste: Effects of pH, temperature, and organic loading rate. Bioresource Technology,
143, 525–530.
Joana Gil-Chávez, G., Villa, J. A., Fernando Ayala-Zavala, J., Basilio Heredia, J., Sepulveda, D.,
Yahia, E. M., & González-Aguilar, G. A. (2013). Technologies for extraction and produc-
tion of bioactive compounds to be used as nutraceuticals and food ingredients: An overview.
Comprehensive Reviews in Food Science and Food Safety, 12(1), 5–23.
Kalogeropoulos, N., Chiou, A., Pyriochou, V., Peristeraki, A., & Karathanos, V. T. (2012).
Bioactive phytochemicals in industrial tomatoes and their processing byproducts. LWT-Food
Science and Technology, 49(2), 213–216.
Karmee, S. K., & Chadha, A. (2005). Preparation of biodiesel from crude oil of Pongamia pinnata.
Bioresource Technology, 96(13), 1425–1429.
Kim, J. K., Oh, B. R., Shin, H. J., Eom, C. Y., & Kim, S. W. (2008). Statistical optimization of
enzymatic saccharification and ethanol fermentation using food waste. Process Biochemistry,
43(11), 1308–1312.
Kumar, K., Yadav, A. N., Kumar, V., Vyas, P., & Dhaliwal, H. S. (2017). Food waste: A poten-
tial bioresource for extraction of nutraceuticals and bioactive compounds. Bioresources and
Bioprocessing, 4, 1–14.
Li, H., Tian, Y., Zuo, W., Zhang, J., Pan, X., Li, L., & Su, X. (2016). Electricity generation
from food wastes and characteristics of organic matters in microbial fuel cell. Bioresource
Technology, 205, 104–110.
Lin, Z., Ooi, J. K., & Woon, K. S. (2022). An integrated life cycle multi-objective optimization
model for health-environment-economic nexus in food waste management sector. Science of
the Total Environment, 816, 151541.
Lipinski, B., Hanson, C., Lomax, J., Kitinoja, L., Waite, R., & Searchinger, T. (2013). Reducing
food loss and waste. Creating a sustainable food future, instalment two. World Resource
Institute.
Memon, M. A. (2010). Integrated solid waste management based on the 3R approach. Journal of
Material Cycles and Waste Management, 12, 30–40.
Muhr, A., Rechberger, E. M., Salerno, A., Reiterer, A., Schiller, M., Kwiecień, M., Adamus, G.,
Kowalczuk, M., Strohmeier, K., Schober, S., & Mittelbach, M. (2013). Biodegradable latexes
from animal-derived waste: Biosynthesis and characterization of mcl-PHA accumulated by Ps.
citronellolis. Reactive and Functional Polymers, 73(10), 1391–1398.
300 A. Sravani et al.

Núñez Selles, A. J., Daglia, M., & Rastrelli, L. (2016). The potential role of mangiferin in can-
cer treatment through its immunomodulatory, anti-angiogenic, apoptopic and gene regulatory
effects. BioFactors, 42(5), 475–491.
Parashar, A., Jin, Y., Mason, B., Chae, M., & Bressler, D. C. (2016). Incorporation of whey per-
meate, a dairy effluent, in ethanol fermentation to provide a zero waste solution for the dairy
industry. Journal of Dairy Science, 99(3), 1859–1867.
Park, J., Lee, B., Tian, D., & Jun, H. (2018). Bioelectrochemical enhancement of methane pro-
duction from highly concentrated food waste in a combined anaerobic digester and microbial
electrolysis cell. Bioresource Technology, 247, 226–233.
Parshetti, G. K., Chowdhury, S., & Balasubramanian, R. (2014). Hydrothermal conversion of
urban food waste to chars for removal of textile dyes from contaminated waters. Bioresource
Technology, 161, 310–319.
Paulraj, C. R. K. J., Bernard, M. A., Raju, J., & Abdulmajid, M. (2019). Sustainable waste manage-
ment through waste to energy technologies in India-opportunities and environmental impacts.
International Journal of Renewable Energy Research (IJRER), 9(1), 309–342.
Pham, T. P. T., Kaushik, R., Parshetti, G. K., Mahmood, R., & Balasubramanian, R. (2015).
Food waste-to-energy conversion technologies: Current status and future directions. Waste
Management, 38, 399–408.
Philippoussis, A. N. (2009). Production of mushrooms using agro-industrial residues as sub-
strates. In Biotechnology for agro-industrial residues utilisation: Utilisation of agro-residues
(pp. 163–196). Springer.
Pimentel, D., & Patzek, T. W. (2005). Ethanol production using corn, switchgrass, and wood;
biodiesel production using soybean and sunflower. Natural Resources Research, 14, 65–76.
Ramírez, I. M., Tsaousi, K., Rudden, M., Marchant, R., Alameda, E. J., Román, M. G., & Banat,
I. M. (2015). Rhamnolipid and surfactin production from olive oil mill waste as sole carbon
source. Bioresource Technology, 198, 231–236.
Reddy, C. S. K., Ghai, R., & Kalia, V. (2003). Polyhydroxyalkanoates: An overview. Bioresource
Technology, 87(2), 137–146.
Riemer, J., & Kristoffersen, M. (1999). Information on waste management practices: A proposed
electronic framework. European Environment Agency.
Rikame, S. S., Mungray, A. A., & Mungray, A. K. (2012). Electricity generation from acido-
genic food waste leachate using dual chamber mediator less microbial fuel cell. International
Biodeterioration & Biodegradation, 75, 131–137.
Roe, B. E., Bender, K., & Qi, D. (2021). The impact of COVID-19 on consumer food waste.
Applied Economic Perspectives and Policy, 43(1), 401–411.
Sadh, P. K., Duhan, S., & Duhan, J. S. (2018). Agro-industrial wastes and their utilization using
solid state fermentation: A review. Bioresources and Bioprocessing, 5(1), 1–15.
Sajna, K. V., Sukumaran, R. K., Jayamurthy, H., Reddy, K. K., Kanjilal, S., Prasad, R. B., &
Pandey, A. (2013). Studies on biosurfactants from Pseudozyma sp. NII 08165 and their poten-
tial application as laundry detergent additives. Biochemical Engineering Journal, 78, 85–92.
Seberini, A. (2020). Economic, social and environmental world impacts of food waste on society
and Zero waste as a global approach to their elimination. In SHS web of conferences (Vol. 74,
p. 03010). EDP Sciences.
Stuart, T. (2009). Waste: Uncovering the global food scandal. WW Norton & Company.
Thyberg, K. L., Tonjes, D. J., & Gurevitch, J. (2015). Quantification of food waste disposal in the
United States: A meta-analysis. Environmental Science & Technology, 49(24), 13946–13953.
Uçkun Kiran, E., Trzcinski, A. P., Ng, W. J., & Liu, Y. (2014). Enzyme production from food
wastes using a biorefinery concept. Waste and Biomass Valorization, 5, 903–917.
Vilas-Boas, A. A., Pintado, M., & Oliveira, A. L. (2021). Natural bioactive compounds from food
waste: Toxicity and safety concerns. Food, 10(7), 1564.
Wu, Y., Ma, H., Zheng, M., & Wang, K. (2015). Lactic acid production from acidogenic fermenta-
tion of fruit and vegetable wastes. Bioresource Technology, 191, 53–58.
Xiong, X., Iris, K. M., Tsang, D. C., Bolan, N. S., Ok, Y. S., Igalavithana, A. D., Kirkham, M. B.,
Kim, K. H., & Vikrant, K. (2019). Value-added chemicals from food supply chain wastes:
State-of-the-art review and future prospects. Chemical Engineering Journal, 375, 121983.
Value-Added Product Development Utilising the Food Wastes 301

Yahia, E. M., & Carrillo-Lopez, A. (Eds.). (2018). Postharvest physiology and biochemistry of
fruits and vegetables. Woodhead Publishing.
Yan, S., Li, J., Chen, X., Wu, J., Wang, P., Ye, J., & Yao, J. (2011). Enzymatical hydrolysis of food
waste and ethanol production from the hydrolysate. Renewable Energy, 36(4), 1259–1265.
Yan, S., Chen, X., Wu, J., & Wang, P. (2013). Pilot-scale production of fuel ethanol from con-
centrated food waste hydrolysates using Saccharomyces cerevisiae H058. Bioprocess and
Biosystems Engineering, 36, 937–946.
Yusoff, M. Z. M., Hu, A., Feng, C., Maeda, T., Shirai, Y., Hassan, M. A., & Yu, C. P. (2013).
Influence of pretreated activated sludge for electricity generation in microbial fuel cell applica-
tion. Bioresource Technology, 145, 90–96.
Zhu, N. M., Luo, T., Guo, X. J., Zhang, H., & Deng, Y. (2015). Nutrition potential of biogas resi-
dues as organic fertilizer regarding the speciation and leachability of inorganic metal elements.
Environmental Technology, 36(8), 992–1000.
Zungur Bastıoglu, A., Tomruk, D., Koç, M., & Ertekin, F. K. (2016). Spray dried melon seed milk
powder: Physical, rheological and sensory properties. Journal of Food Science and Technology,
53, 2396–2404.
Role of Bacterial Degradation
in Lignocellulosic Biomass for Biofuel
Production

Arti Kumari, Maneesh Kumar, and Bibekananda Bhoi

1 Introduction

Lignocellulosic biomass (LCB), the most abundant and renewable natural resource
on Earth, offers significant economic and strategic advantages for biofuel produc-
tion. It is mainly composed of cellulose, hemicellulose, and lignin and can be a
cost-effective source of fermentable sugars. Its use in terms of green, renewable
energy reduces the dependency on fossil fuels (Patel & Shah, 2021). However, lig-
nin presents a challenge because it forms a physical barrier and inhibits the hydro-
lysis of cellulose and hemicellulose. Physicochemical pretreatment methods have
been developed to improve accessibility, but they are costly and energy-intensive.
Enzymatic hydrolysis offers a more specific and environmentally friendly approach
(Haldar et al., 2022). While fungi are known to produce lignocellulose-degrading
enzymes, their use is time-consuming and expensive. According to studies, numer-
ous microorganisms can contribute to the degradation of lignocellulosic biomass,
and numerous fungal and bacterial species have been shown to produce cellulolytic
and hemicellulolytic enzymes in both aerobic and anaerobic environments.
Therefore, the discovery of new lignocellulose-degrading bacteria is critical for bio-
fuel production and other industrial applications. Compared to other sources,
enzymes from microbial sources are preferred because they are easier to culture and
modify for desired yields. The most widespread biological organisms in nature are

A. Kumari (*)
Department of Biotechnology, Patna Women’s College, Patna, Bihar, India
e-mail: arti.mbio@patnawomenscollege.in
M. Kumar
Department of Biotechnology, Magadh University, Bodh Gaya, Bihar, India
B. Bhoi
Department of Microbiology, Dr. Harisingh Gour Vishwavidyalaya (A Central University),
Sagar, Madhya Pradesh, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 303
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_17
304 A. Kumari et al.

probably bacterial and fungal species that can degrade both natural and synthetic
polymers (Mathews et al., 2019).
Microbial degradation of lignocellulosic biomass (LCB) produces important
products such as 2G-ethanol (Banu et al., 2020), in which humans are interested.
Degradation of lignocellulose requires microbial enzymes such as glycosyl hydro-
lases, carbohydrate esterases, and auxiliary enzymes. LCB produced from abundant
and cheap agricultural wastes, grasses, waste paper, food waste, and municipal solid
waste is a promising source of renewable energy, especially liquid fuels (Al-Battashi
et al., 2019). It is mainly composed of cellulose, hemicellulose, and lignin, but the
distribution and structure of plants and plant parts vary. LCB is a promising method
for biofuel production that does not compete with food production (Savla
et al., 2021).

2 Significance of Bacteria in Biomass Production

Fungi, especially white rot, degrade lignin, which has been extensively studied
(Sista Kameshwar & Qin, 2018). Fungi are the most important decomposers,
although their genetic engineering potential is still low compared to other species.
Commercially produced fungal enzymes are expensive and unselective. They have
a hard time in difficult situations because they cannot withstand or adapt to chang-
ing environmental conditions (Jones et al., 2018; Li et al., 2023). Bacteria and their
enzymes are more adaptable to pH and temperature than fungi. Recent research on
synthetic microbial communities has shown that bacteria influence the enzyme
activity of lignocellulose more than fungi (Hu et al., 2017; Li et al., 2023). To under-
stand their biological significance and function, model organisms for bacterial gluc-
uronoyl esterases (GEs) from the carbohydrate esterase family 15 were biochemically
characterized and structurally identified. Only a few enzymes surpassed the previ-
ously known fungal GEs in their catalytic efficiency (Arnling Bååth et al., 2018). A
salt marsh soil microbiome-derived, halo-tolerant, lignocellulose-degrading micro-
bial consortium was formed using wheat straw as carbon and energy. Under saline
conditions, bacteria degrade resistant lignocellulose better than fungi, according to
the consortium. The final consortium had larger levels of lignocellulolytic haloen-
zymes than the prior one, supporting the concept that bacteria degrade lignocellu-
lose in salty conditions more than fungi (Sousa et al., 2022).
Bacteria can convert lignocellulose into useful chemicals; therefore, they have
attracted scientific attention due to their versatility (Cortes-Tolalpa et al., 2018). We
need new enzyme combinations, microbial consortia, and bioengineering to improve
promising strains and create microbial communities that can degrade biomass and
produce more biofuel (Chukwuma et al., 2021). Unexpectedly, reducing simple car-
bon sources leads to an increase in bacterial populations, while leaving complex
bacteria and more lignin. Functional diversity, a wide range of terminal electron
Role of Bacterial Degradation in Lignocellulosic Biomass for Biofuel Production 305

acceptors, and the ability to degrade lignin are increasing interest in bacteria for new
biotechnological processes (Ayodeji et al., 2023). Numerous studies using insect
stomachs have demonstrated lignin-degrading bacteria such as Alpha Proteobacteria,
Gamma Proteobacteria, and Actinomycetia (Fig. 1). Industrial production of next-­
generation biofuels requires the development of novel bacterial lignin-degrading
enzymes (Madhavan et al., 2017; Duran et al., 2022).
This is because microorganisms can be engineered for biofuel production, adapt
to oxygen demand, and withstand extreme environments (Antar et al., 2021).
According to a hydrolysis study, aerobic bacteria most commonly use the free
enzyme system to digest lignocellulosic biomass. Anaerobic bacteria use cellulo-
somes and xylanosomes, complex protein structures, as supporting enzymes to
degrade biomass (Singh et al., 2021). Bacteria can also synthesize hydrolytic
enzymes needed to degrade lignocellulosic biomass (Rai et al., 2019a, b).
Acetovibrio, Bacillus, Bacteroides, Cellulomanas, Clostridium, Erwinia,
Microbispora, Ruminococcus, Streptomyces, and Thermomonospora degrade ligno-
cellulosic biomass according to various studies (Chukwuma et al., 2021).
Acetovibrio and Ruminococcus, for example, are known for their ability to
degrade cellulose in the rumen of animals. Bacillus and Clostridium are well-­studied
cellulolytic bacteria, and Cellulomonas is known for its cellulase production.
Streptomyces, while primarily known for antibiotic production, also contributes to
the degradation of lignocellulose. These bacteria play a critical role in nature’s recy-
cling process by breaking down lignocellulosic biomass into simpler sugars and
organic compounds. This process is essential for nutrient cycling and contributes to
ecosystem sustainability (Arnling Bååth et al., 2018). Understanding the specific
role of these bacteria in the degradation of lignocellulose is crucial not only for
environmental studies, but also for the development of biotechnological applica-
tions such as biofuel production and waste management strategies that take advan-
tage of the natural capabilities of these bacteria (Chukwuma et al., 2021).

Pretreatment of
Enzymatic Hydrolysis
lignocellulosic Biomass
(Cellulases, Sugars & Lignin
(physical, chemical &
Hemicellulses)
biological

High Value Bioproducts


(Enzymes, Oligomers,
Lignin compounds, Fermentation of sugars
Glycerol, Xylitol, Acetoin
& 2,3-butanediol

Fig. 1 Steps involved in biotechnological conversion of lignocellulosic biomass


306 A. Kumari et al.

3 Lignocellulose-Degrading Bacteria

Lignocellulose-degrading bacteria use many important methods to degrade plant


cell walls. The production of cellulases, hemicellulases, and ligninases is essential
(Silva et al., 2018). These enzymes synergistically hydrolyze cellulose and hemicel-
lulose to release sugars for microbes. These bacteria have specialized surface adhe-
sion mechanisms to attach to lignocellulosic substrates and facilitate enzymatic
degradation (Wu et al., 2022). Some strains degrade lignin via extracellular redox
systems. These methods allow lignocellulosic-degrading bacteria to efficiently
degrade plant biomass into simpler chemicals, which improves nutrient cycling,
biofuel production, and waste management (Rout et al., 2022). Microbes that con-
vert lignocellulose into compounds are being increasingly studied. Their excellent
suitability and versatility in this method are particularly important (Moreno et al.,
2020). Bacterial communities survive when basic carbon sources are scarce, while
more complicated communities thrive at higher lignin concentrations. The func-
tional diversity of bacteria, their wide range of terminal electron acceptors, and their
lignin degradation make them biotechnological options (Chukwuma et al., 2021).
Most bacteria degrade lignocellulosic biomass with free enzymes. Anaerobic bacte-
ria support biomass hydrolysis enzymes with cellulosomes or xylanosomes. Finally,
multicatalytic enzyme systems that combine numerous effects in one gene product
are rarely studied. These structures help biomass hydrolysis enzymes. Multicatalytic
enzyme systems combine multiple capabilities in one gene (Champreda et al., 2019).

4 Cellulose Degradation

Lignocellulose is mostly broken down by cellulolytic bacteria, which are mostly


found in the order Actinomycetales (phylum Actinobacteria) and the anaerobic
order Clostridiales (phylum Firmicutes). Both the aerobic Actinomycetales and the
anaerobic Clostridiales are capable of degrading cellulose. Anaerobic clostridia
multiply on the surface of cellulosic material and release complex cellulases that
degrade cellulose since they cannot penetrate the material. Aerobic bacteria use free
cellulolytic enzymes to degrade the cellulose in the biomass (Villota et al., 2020). In
this process, cellulose is first hydrolyzed to cellobiose, which is then fermented to
produce carbon dioxide, hydrogen, and organic acids (Table 1). The predominant
bacteria use these secondary products to produce various valuable compounds
(Deng, Kumar, and Wang, 2014; Robak & Balcerek, 2018). Table 1 shows the types
of bacteria that can be useful in degrading lignocellulose.
In contrast, bacteria use metabolic pathways to degrade organic material in
anaerobic biodegradation. Bacterial anaerobic digestion of biomass converts sugars
into alcohols or acids, producing biogas. Primary microorganisms, or cellulolytic
microbes, degrade cellulose and produce sugars for growth and maintenance.
Secondary microorganisms, which cannot hydrolyze cellulose, survive on main
Role of Bacterial Degradation in Lignocellulosic Biomass for Biofuel Production 307

Table 1 List of lignocellulytic bacteria


Sl.
no Lignocellulosic substrate Degrading bacteria References
1. Wheat straw, willow, oak Bacillus subtilis, B. pumilus, Acharya and Chaudhary (2012)
wood B. licheniformis
2. Palm empty fruit bunch, B. cereus, B. Yadav et al. (2020)
oak wood, eucalyptus amyloliquefaciens
wood, maize cob
3. Pineapple leaves Paenibacillus sp. Vaithanomsat et al. (2022)
4. Switchgrass Ruminococcus albus Moraïs and Mizrahi (2019)
5. Corn stover Cellulomonas fimi Batool et al. (2018)
6. Sugarcane bagasse Clostridium thermocellum An et al. (2020)
7. Miscanthus Caldicellulosiruptor sp. Svetlitchnyi et al. (2022)
8. Cotton stalk Erwinia chrysanthemi Yudden et al. (2019)
9. Bagasse Cellulomonas uda Swathy et al. (2020)
10. Poplar wood Pseudomonas fluorescens Liu et al. (2019)
11. Spruce wood Streptomyces griseus Toussaint et al. (2016)
12. Olive pomace Aspergillus niger Oliveira et al. (2016)
13. Sunflower stalk Trichoderma koningii El-Ibrahime and Mourad (2020)
14. Spruce wood Trichoderma hamatum Pusz and Zwijacz-­Kozica (2017)
15. Potato pulp Aspergillus oryzae Zuo et al. (2018)

microbe products. Symbiotic connections allow secondary bacteria to ingest free


sugars from primary microbes, degrading cellulose. This “symbiosis” is essential
for cellulose breakdown. In aerobic conditions, primary and secondary microorgan-
isms work together to completely oxidize cellulose to CO2 (Moraïs & Mizrahi, 2019).

5 Lignin Degradation

The presence of lignin, the second most abundant biomacromolecule after cellulose,
complicates the degradation of lignocellulose. Lignin is a complex phenolic macro-
molecule found in thicker secondary cell walls of plants. It structures the plant,
makes the cell walls impermeable, and protects them from microbes and enzymes.
Cross-linked polymers of coniferyl, coumaryl, and sinapyl alcohols—guaiacyl (G),
syringyl (S), and p-hydroxyphenyl (H) phenylpropanoid units—make it. These
monoolignols are linked by alkyl-aryl, alkyl-alkyl, and aryl-aryl ethers in the lignin
polymer. The monoolignol mix depends on the lignin source. Angiosperms have
p-coumaryl alcohol, guaiacyl units, and syringyl units, while hardwoods have sev-
eral guaiacyl units (Hapuarachchi et al., 2021).
Streptomyces “Actinomycetales” has long been known for delignification.
Ligninolytic bacteria that degrade lignin are widely distributed in nature. Another
known microbe, Clostridium thermocellum, uses sugars from lignin degradation for
energy production. Nutrient cycling improves soil bioequilibrium (De Bhowmick
308 A. Kumari et al.

et al., 2018). Tunnelling, erosion, and cavitation help bacteria break down lignified
cell walls. Bacteria such as Actinobacteria and Proteobacteria depolymerize lignin,
degrade aromatic compounds and biosynthesize certain chemicals (Zuliani et al.,
2021). Streptomyces, Sphingomonas, Pseudomonas, Rhodococcus, and Nocardia
have been extensively studied for lignin degradation. Research shows that bacteria
degrade lignin through various mechanisms and produce useful bioproducts.
Cellulose is converted in the rumen to important animal acids, including acetate,
butyrate, and propionate. In digesters and under other anaerobic conditions, CH4
and CO2 are produced. The first stage of anaerobic degradation of cellulose requires
H2-producing bacteria, while the second stage requires H2-consuming bacteria.
Cellulolytic bacteria such as Clostridium degrade lignocellulose anaerobically.
Multienzyme complexes and cellulase enzyme synthesis are the specialty of these
bacteria. Other cellulolytic groups include Acetovibrio, Bacillus, Bacteriodes,
Cellulomonas, Ruminococcus, Thermomonospora, Erwinia, and Microbispora
(Srivastava et al., 2019).

6 Dual Degradation Approach of Cellulose and Lignin

The new “dual degradation” method for the degradation of lignocellulose improves
the efficiency of biomass conversion by simultaneously degrading cellulose and
lignin. This method takes into account the complicated structure of plant cell walls
composed of cross-linked cellulose, hemicellulose, and lignin (Zeng et al., 2021).
This method selectively degrades cellulose and lignin, which holds promise for bio-
fuel production, bioremediation, and sustainable waste management. Microbes
have been shown to be critical for both phases of this degradation process (Sharma
et al., 2019). Fungi and bacteria with lignocellulolytic capabilities can produce a
variety of enzymes that degrade cellulose and lignin synergistically. Microbes
metabolize cellulose and hemicellulose after cellulases and hemicellulases hydro-
lyze them to simpler sugars. Ligninases or eroxidases modify lignin and degrade it,
making cellulose more enzymatically active (Henske et al., 2018).
Powerful lignin-modifying enzymes from the white-rot fungus are required for
this dual degradation process. Lignin peroxidases, manganese peroxidases, and lac-
cases degrade lignin and expose cellulose (Lankiewicz et al., 2022). Some
Actinobacteria and Firmicutes strains produce cellulolytic and ligninolytic enzymes
that cause dual degradation. The ability to microbially degrade lignocellulose is
enhanced by the presence of fungi in combination with their flexibility, either inde-
pendently or in conjunction with each other. Dual degradation threatens the sustain-
ability of biofuels. By degrading cellulose to fermentable sugars and reducing
lignin, biomass can be more effectively converted to biofuels such as ethanol. This
approach enhances and cleans the environment by reducing the refractoriness of
lignocellulosic waste. These bacteria act synergistically in bioenergy, bioremedia-
tion, and sustainable waste management (Hussain et al., 2023).
Role of Bacterial Degradation in Lignocellulosic Biomass for Biofuel Production 309

7 Bacteria Involved in Degradation


of Lignocellulosic Biomass

Lignocellulosic biomass is a good carbon source for sugars and other chemicals.
Due to the complexity and diversity of the material, multiple hydrolytic enzymes
are required for biodegradation. Complex microbial populations interact regularly
in nature to degrade. Investigating synergies in LCB degradation is critical for opti-
mizing biodegradation. We tested artificial microbial consortia of bacteria and fungi
to degrade wheat straw (Rajeswari et al., 2021). Wheat straw was the sole carbon
and energy source for the aerobic cultures used to study the development of mono-
cultures of degrading strains and the release of enzymes. To test synergism, selected
strains were grown in co-culture and compared with monocultures. Organisms in
monoculture mainly consume cellulose (Sagarika et al., 2022). Each organism has
a unique enzyme profile. One strain, Flavobacterium ginsengisoli so9, had a strong
enzyme release and degradation capacity. Five of 13 co-cultures acted synergisti-
cally. Four bacterial bicultures and one bacterial–fungal triculture were included.
Citrobacter freundii and Sphingobacterium multivorum produced 18.2 times more
biomass in a biculture, showing the highest synergy (Karnaouri et al., 2022). This
bacterial pair showed stronger enzymatic activity than the two monocultures, espe-
cially cellobiohydrolases, mannosidases, and xylosidases. Synergistic effects
occurred only in wheat straw bicultures and were not observed in glucose bicul-
tures. It was found that the wasted supernatants of both partners specifically
enhanced the development of the wheat straw counterpart. Chemicals produced by
S. multivorum w15 can enhance C. freundii so4 activity and vice versa, resulting in
LCB-specific synergism (Ilić et al., 2023).
Bacterial systems are attractive for depolymerization of lignocellulosic biomass
due to their functional diversity and adaptability. To optimize microbial utilization
of lignocellulose, multiple metabolic pathways and enzymes with different speci-
ficities are required to efficiently utilize lignocellulosic waste and produce bioprod-
ucts in biorefineries (Ashokkumar et al., 2022). This study investigated the metabolic
flexibility and enzymatic capacity of highly diverse aerobic mesophilic bacteria to
lignocellulosic substrates. Lake Keri, a pristine wetland with biomass loss and sub-
surface oil leaks, supported these microorganisms (Vaithanomsat et al., 2022). In
enrichment cultures with Organosolv lignin as the sole carbon and energy source,
many Pseudomonas bacteria absorbed lignin-associated aromatic compounds.
Actinobacteria, Proteobacteria, Sphingobacteria, and Flavobacteria were also
more abundant in cultures grown exclusively on xylan or carboxymethylcellulose
(Ghosh et al., 2017). Several different individuals could target structural lignocel-
lulose polysaccharides by hydrolyzing crystalline or amorphous cellulose and
xylan. Lignin hydrolysates from agricultural waste pretreated with alkali improved
the growth ability of certain isolates. The keri isolates are promising for commercial
use in biorefineries as lignocellulose degraders (Karnaouri et al., 2022).
Bioengineering can improve promising strains and build biomass-degrading
microbial communities (Prasad et al., 2019). Novel enzyme combinations and
310 A. Kumari et al.

microbial consortia are needed to degrade lignocellulosic biomass for biofuel pro-
duction. Bacterial cooperation is essential for lignocellulosic biomass degradation
(Sharma et al., 2019). The degradation of lignocellulose, hemicellulose, and lignin
requires multistage bacterial cooperation. Wheat straw, corn stover, sugarcane
bagasse, and other agricultural wastes can be used to produce lignocellulose, which
is a promising feedstock for the production of valuable raw materials and fuels (Ilić
et al., 2023). However, its complex chemical composition and physical structure
make LCB difficult to degrade. In wheat straw, LCB resistance is due to the tight
bonds between lignin, hemicellulose, and cellulose, as well as the degree of crystal-
lization and polymerization. Synergies between degrading organisms are common
in microbial communities and are critical for LCB degradation. Synergistic growth
occurs when an assemblage of microorganisms produces more biomass on a sub-
strate than expected. In contrast, “enzymatic synergism” describes the cooperation
of organisms that use enzymes with different but overlapping tasks. These two syn-
ergies are associated in LCB-degrading microbial populations (Swathy et al., 2020).
Niche partitioning and metabolic complementarity provide synergy in LCB deg-
radation. Microorganisms partition niches by using LCB components in multiple
metabolic pathways. Cultures share metabolism. Co-cultures of Trichoderma reesei
and E. coli produced isobutanol from corn stover. The cellulolytic enzymes of
T. reesei converted LCB to soluble saccharides, which E. coli fermented to isobuta-
nol. Substrate composition and structure influence the interactions between the deg-
radation products. The behaviors of the cooperating animals are unknown
(Ashokkumar et al., 2022). Microbial attack on lignocellulose and substrate compo-
sition and structure in the dynamics of cooperating organisms need further study.
Strains of lignocellulolytic microbial consortia were selected by repeated growth on
raw wheat straw. It was hypothesized that the diverse and spatially structured wheat
straw substrate would allow lignocellulolytic bacteria to cooperate. Synergies were
investigated in the study. De Bhowmick et al. (2018) found cooperative wheat straw
microbial consortia, but no glucose.

8 Enzymes for Biodegradation of Lignocellulose

Enzymes simplify biomolecules such as lignocellulose for commercial and biotech-


nological use. Lignocellulases are essential in lignocellulosic biorefineries.
Lignocellulases include amylases, proteases, esterases, hemicellulases, pectinases,
and chitinases. Their low-cost, eco-friendly, and energy-efficient production has
attracted much attention (Lankiewicz et al., 2022). The global enzymes market
includes lignocellulases due to their increasing use in industrial and biotechnologi-
cal applications. CAZymes produce, modify, and cleave carbohydrates. Glycosyl
hydrolases (GH), polysaccharide lyases (PL), carbohydrate esterases (CE), and gly-
cosyltransferases (GT) are classified based on amino acid sequence, enzyme func-
tion, and protein folding. Glycosyl hydrolases efficiently convert lignocellulosic
Role of Bacterial Degradation in Lignocellulosic Biomass for Biofuel Production 311

biomass (LCB) into fermentable sugars. These enzymes break glycosidic oligosac-
charide bonds (Liu et al., 2019).
Amylases, (hemi-) cellulases, chitinases, and pectinases are lignocellulolytic
enzymes used in manufacturing and biorefining. Extracellular enzymes from bacte-
ria that can withstand a wide range of temperatures and pH are suitable for these
purposes. Extracellular enzymes are produced by many bacteria. These include
Pseudomonas, Enterobacter, Bacillus, Klebsiella, Paenibacillus, Rhodococcus,
Cellulomonas, Streptomyces, and Citrobacter. The Bacillus species are heat resis-
tant. They produce lignocellulases and are thus lignocellulolytic. According to Sar
et al. (2023), B. subtilis, which produces the most lignin peroxides (LiP) and cellu-
lases, can compost rice straw. Another study investigated the ability of Bacillus sp.
R2 to produce lignocellulolytic enzymes from coffee waste after various pretreat-
ments. Extremophilic enzymes thrive in extreme conditions, making them useful to
industry. Both laccase and peroxidase are catalytic oxidative enzymes produced by
actinobacteria (Kumar et al., 2021). Caldicellulosiruptor species produce glycoside
hydrolases (GH) that degrade cellulose at temperatures up to 70 °C. In laboratory
tests, GH combinations of three Caldicellulosiruptor species degraded microcrys-
talline cellulose better than single enzyme. Classification and synthesis of bacterial
lignocellulases could lead to lignocellulosic biorefineries. As their enzymatic capa-
bilities are better known, Bacillus, Actinobacteria, and Caldicellulosiruptor species
can be used in biomass hydrolysis and bioenergy production (Chukwuma et al.,
2021). Biochemical enzymes are an environmentally friendly and efficient alterna-
tive to standard industrial processes. Biological enzymes consume less energy and
have a lower environmental impact than chemical processes. The biorefinery indus-
try is booming, and more environmentally friendly energy and chemical production
is desired, increasing the demand for lignocellulolytic enzymes (Ashokkumar
et al., 2022).

9 Conclusion

Bacterial bioconversion of lignocellulosic biomass offers much untapped potential,


but more studies need to be done to uncover these opportunities and figure out how
to exploit them. Although genetic engineering and the development of “omics”
technologies hold promise for finding new bacterial communities that have not yet
been studied for their ability to degrade lignocellulosic materials, these methods are
lengthy and costly. Molecular-level studies do not eliminate the need for culture-­
dependent methods to study the optimal functionality of an organism to facilitate its
incorporation into industrial processes. It is necessary to conduct additional studies
to find adaptable and unique bacterial species that can be explored and manipulated
to achieve a more environmentally friendly outcome. Bacteria with multienzyme
activities and the ability to survive in extreme environments will play an important
role in the production cycle because they are more versatile and can perform
312 A. Kumari et al.

multiple functions in the cycle. Collaborations and conditions conducive to research


are critical if the use of bacteria in lignocellulosic biorefineries is to increase so that
the associated benefits can be realized.

References

Acharya, S., & Chaudhary, A. (2012). Optimization of fermentation conditions for cellulases
production by Bacillus licheniformis MVS1 and Bacillus sp. MVS3 isolated from Indian hot
spring. Brazilian Archives of Biology and Technology, 55, 497–503.
Al-Battashi, H. S., Annamalai, N., Sivakumar, N., Al-Bahry, S., Tripathi, B. N., Nguyen, Q. D.,
& Gupta, V. K. (2019). Lignocellulosic biomass (LCB): a potential alternative biorefinery
feedstock for polyhydroxyalkanoates production. Reviews in Environmental Science and Bio/
Technology, 18, 183–205.
An, Q., Bu, J., Cheng, J. R., Hu, B. B., Wang, Y. T., & Zhu, M. J. (2020). Biological saccharification
by Clostridium thermocellum and two-stage hydrogen and methane production from hydrogen
peroxide-acetic acid pretreated sugarcane bagasse. International Journal of Hydrogen Energy,
45(55), 30211–30221.
Antar, M., Lyu, D., Nazari, M., Shah, A., Zhou, X., & Smith, D. L. (2021). Biomass for a sustain-
able bioeconomy: An overview of world biomass production and utilization. Renewable and
Sustainable Energy Reviews, 139, 110691.
Arnling Bååth, J., Mazurkewich, S., Knudsen, R. M., Poulsen, J. C. N., Olsson, L., Lo Leggio, L.,
& Larsbrink, J. (2018). Biochemical and structural features of diverse bacterial glucuronoyl
esterases facilitating recalcitrant biomass conversion. Biotechnology for Biofuels, 11(1), 1–14.
Ashokkumar, V., Venkatkarthick, R., Jayashree, S., Chuetor, S., Dharmaraj, S., Kumar, G., Chen,
W. H., & Ngamcharussrivichai, C. (2022). Recent advances in lignocellulosic biomass for bio-
fuels and value-added bioproducts-A critical review. Bioresource Technology, 344, 126195.
Ayodeji, F. D., Shava, B., Iqbal, H. M., Ashraf, S. S., Cui, J., Franco, M., & Bilal, M. (2023).
Biocatalytic versatilities and biotechnological prospects of laccase for a sustainable industry.
Catalysis Letters, 153(7), 1932–1956.
Banu, J. R., Merrylin, J., Usman, T. M., Kannah, R. Y., Gunasekaran, M., Kim, S. H., & Kumar,
G. (2020). Impact of pretreatment on food waste for biohydrogen production: A review.
International Journal of Hydrogen Energy, 45(36), 18211–18225.
Batool, I., Gulfraz, M., Asad, M. J., Kabir, F., Khadam, S., & Ahmed, A. (2018). Cellulomonas
sp. isolated from termite gut for saccharification and fermentation of agricultural biomass.
BioResources, 13(1), 752–763.
Champreda, V., Mhuantong, W., Lekakarn, H., Bunterngsook, B., Kanokratana, P., Zhao, X. Q.,
Zhang, F., Inoue, H., Fujii, T., & Eurwilaichitr, L. (2019). Designing cellulolytic enzyme sys-
tems for biorefinery: from nature to application. Journal of bioscience and bioengineering,
128(6), 637–654.
Chukwuma, O. B., Rafatullah, M., Tajarudin, H. A., & Ismail, N. (2021). A review on bacterial
contribution to lignocellulose breakdown into useful bio-products. International journal of
environmental research and public health, 18(11), 6001.
Cortes-Tolalpa, L., Norder, J., van Elsas, J. D., & Falcao Salles, J. (2018). Halotolerant microbial
consortia able to degrade highly recalcitrant plant biomass substrate. Applied Microbiology and
Biotechnology, 102, 2913–2927.
De Bhowmick, G., Sarmah, A. K., & Sen, R. (2018). Lignocellulosic biorefinery as a model for
sustainable development of biofuels and value added products. Bioresource Technology, 247,
1144–1154.
Deng, Y., Kumar, S., & Wang, H. (2014). Synergistic–cooperative combination of enamine cataly-
sis with transition metal catalysis. Chemical Communications, 50(33), 4272–4284.
Role of Bacterial Degradation in Lignocellulosic Biomass for Biofuel Production 313

Duran, K., van den Dikkenberg, M., van Erven, G., Baars, J. J., Comans, R. N., Kuyper, T. W., &
Kabel, M. A. (2022). Microbial lignin degradation in an industrial composting environment.
Bioresource Technology Reports, 17, 100911.
El-Ibrahime, I. A., & Mourad, K. A. (2020). Efficacy of some Trichoderma species on manage-
ment of sunflower head-rot. Journal of Plant Protection and Pathology, 11(11), 537–542.
Ghosh, K., Banerjee, S., Moon, U . M., Khan, H. A., & Dutta, D. (2017). Evaluation of gut associ-
ated extracellular enzyme-producing and pathogen inhibitory microbial community as poten-
tial probiotics in Nile tilapia, Oreochromis niloticus. International Journal of Aquaculture, 7.
Haldar, D., Dey, P., Patel, A. K., Dong, C. D., & Singhania, R. R. (2022). A critical review on the
effect of lignin redeposition on biomass in controlling the process of enzymatic hydrolysis.
Bioenergy Research, 15, 1–12.
Hapuarachchi, K. K., Karunarathna, S. C., Xu, X. H., Dutta, A. K., Phengsintham, P., Hyde, K. D.,
& Wen, T. C. (2021). A review on bioactive compounds, beneficial properties and biotech-
nological approaches of Trametes (Polyporaceae, Polyporales) and a new record from Laos.
Chiang Mai Journal of Science, 48(3), 674–698.
Henske, J. K., Wilken, S. E., Solomon, K. V., Smallwood, C. R., Shutthanandan, V., Evans,
J. E., Theodorou, M. K., & O'Malley, M. A. (2018). Metabolic characterization of anaerobic
fungi provides a path forward for bioprocessing of crude lignocellulose. Biotechnology and
Bioengineering, 115(4), 874–884.
Hu, J., Xue, Y., Guo, H., Gao, M. T., Li, J., Zhang, S., & Tsang, Y. F. (2017). Design and composi-
tion of synthetic fungal-bacterial microbial consortia that improve lignocellulolytic enzyme
activity. Bioresource Technology, 227, 247–255.
Hussain, C. M., Bharagava, R. N., Goswami, L., & Kushwaha, A. (Eds.). (2023). Bio-based mate-
rials and waste for energy generation and resource management: Volume 5 of Advanced zero
waste tools: Present and emerging waste management practices. Elsevier.
Ilić, N., Milić, M., Beluhan, S., & Dimitrijević-Branković, S. (2023). Cellulases: From lignocel-
lulosic biomass to improved production. Energies, 16(8), 3598.
Jones, J. A., Kerr, R. G., Haltli, B. A., & Tinto, W. F. (2018). Temperature and pH effect on glu-
cose production from pretreated bagasse by a novel species of Citrobacter and other bacteria.
Heliyon, 4(6), e00657.
Karnaouri, A., Chorozian, K., Zouraris, D., Karantonis, A., Topakas, E., Rova, U., &
Christakopoulos, P. (2022). Lytic polysaccharide monooxygenases as powerful tools in
enzymatically assisted preparation of nano-scaled cellulose from lignocellulose: A review.
Bioresource Technology, 345, 126491.
Kumar, A., Singh, A. K., Bilal, M., & Chandra, R. (2021). Extremophilic ligninolytic enzymes:
Versatile biocatalytic tools with impressive biotechnological potential. Catalysis Letters,
152, 1–25.
Lankiewicz, T. S., Lillington, S. P., & O’Malley, M. A. (2022). Enzyme discovery in anaerobic
fungi (Neocallimastigomycetes) enables lignocellulosic biorefinery innovation. Microbiology
and Molecular Biology Reviews, 86(4), e00041-22.
Li, W., Ayers, C., Huang, W., Schilling, J. S., Cullen, D., & Zhang, J. (2023). A laccase gene
reporting system that enables genetic manipulations in a brown rot wood decomposer fungus
Gloeophyllum trabeum. Microbiology Spectrum, 11(1), e04246-22.
Liu, H., Wu, X., Ren, J., & Chen, D. (2019). Effect of co-inoculation with Pseudomonas fluores-
cens and Xerocomus chrysenteron on the soil enzyme activity and microbial diversity in poplar
rhizosphere. Scientia Silvae Sinicae, 55(1), 22–30.
Madhavan, A., Sindhu, R., Parameswaran, B., Sukumaran, R. K., & Pandey, A. (2017).
Metagenome analysis: A powerful tool for enzyme bioprospecting. Applied Biochemistry and
Biotechnology, 183, 636–651.
Mathews, S. L., Epps, M. J., Blackburn, R. K., Goshe, M. B., Grunden, A. M., & Dunn, R. R. (2019).
Public questions spur the discovery of new bacterial species associated with lignin bioconver-
sion of industrial waste. Royal Society Open Science, 6(3), 180748.
Moraïs, S., & Mizrahi, I. (2019). Islands in the stream: From individual to communal fiber degra-
dation in the rumen ecosystem. FEMS Microbiology Reviews, 43(4), 362–379.
314 A. Kumari et al.

Moreno, A. D., Ibarra, D., Eugenio, M. E., & Tomás-Pejó, E. (2020). Laccases as versatile enzymes:
From industrial uses to novel applications. Journal of Chemical Technology & Biotechnology,
95(3), 481–494.
Oliveira, F., Moreira, C., Salgado, J. M., Abrunhosa, L., Venâncio, A., & Belo, I. (2016). Olive
pomace valorization by Aspergillus species: Lipase production using solid-state fermentation.
Journal of the Science of Food and Agriculture, 96(10), 3583–3589.
Patel, A., & Shah, A. R. (2021). Integrated lignocellulosic biorefinery: Gateway for production of
second generation ethanol and value added products. Journal of Bioresources and Bioproducts,
6(2), 108–128.
Prasad, R. K., Chatterjee, S., Mazumder, P. B., Gupta, S. K., Sharma, S., Vairale, M. G., Datta, S.,
Dwivedi, S. K., & Gupta, D. K. (2019). Bioethanol production from waste lignocelluloses: A
review on microbial degradation potential. Chemosphere, 231, 588–606.
Pusz, W., & Zwijacz-Kozica, T. (2017). Composition of microscopic fungi associated with the
spruce dead wood in the Tatra National Park. Sylwan, 161(4), 312–319.
Rai, R., Agrawal, D., & Chadha, B. S. (2019a). New paradigm in degradation of lignocellulosic
biomass and discovery of novel microbial strains. In Microbial diversity in ecosystem sustain-
ability and biotechnological applications: Volume 2. Soil & Agroecosystems (pp. 403–440).
Springer.
Rai, M., Ingle, A. P., Pandit, R., Paralikar, P., Biswas, J. K., & da Silva, S. S. (2019b). Emerging
role of nanobiocatalysts in hydrolysis of lignocellulosic biomass leading to sustainable bio-
ethanol production. Catalysis Reviews, 61(1), 1–26.
Rajeswari, G., Jacob, S., Chandel, A. K., & Kumar, V. (2021). Unlocking the potential of insect and
ruminant host symbionts for recycling of lignocellulosic carbon with a biorefinery approach: A
review. Microbial Cell Factories, 20(1), 107.
Robak, K., & Balcerek, M. (2018). Review of second generation bioethanol production from resid-
ual biomass. Food Technology and Biotechnology, 56(2), 174.
Rout, M., Sardar, B., Singh, P. K., Pattnaik, R., & Mishra, S. (2022). Utilization of microbial
potential for bioethanol production from lignocellulosic waste. In Biotechnology for zero
waste: Emerging waste management techniques (pp. 263–282). Wiley-VCH.
Sagarika, M. S., Parameswaran, C., Senapati, A., Barala, J., Mitra, D., Prabhukarthikeyan, S. R.,
Kumar, A., Nayak, A. K., & Panneerselvam, P. (2022). Lytic polysaccharide monooxygenases
(LPMOs) producing microbes: A novel approach for rapid recycling of agricultural wastes.
Science of the Total Environment, 806, 150451.
Sar, T., Harirchi, S., Yazdian, F., & Taherzadeh, M. J. (2023). Microorganisms used. In Biotic
resources: Circular bioeconomy perspective (p. 21). CRC Press.
Savla, N., Pandit, S., Mathuriya, A. S., Gupta, P. K., Khanna, N., Babu, R. P., & Kumar, S. (2021).
Recent advances in bioelectricity generation through the simultaneous valorization of lig-
nocellulosic biomass and wastewater treatment in microbial fuel cell. Sustainable Energy
Technologies and Assessments, 48, 101572.
Sharma, H. K., Xu, C., & Qin, W. (2019). Biological pretreatment of lignocellulosic biomass for
biofuels and bioproducts: An overview. Waste and Biomass Valorization, 10, 235–251.
Silva, C. O., Vaz, R. P., & Filho, E. X. (2018). Bringing plant cell wall-degrading enzymes into
the lignocellulosic biorefinery concept. Biofuels, Bioproducts and Biorefining, 12(2), 277–289.
Singh, N., Mathur, A. S., Gupta, R. P., Barrow, C. J., Tuli, D. K., & Puri, M. (2021). Enzyme sys-
tems of thermophilic anaerobic bacteria for lignocellulosic biomass conversion. International
Journal of Biological Macromolecules, 168, 572–590.
Sista Kameshwar, A. K., & Qin, W. (2018). Comparative study of genome-wide plant biomass-­
degrading CAZymes in white rot, brown rot and soft rot fungi. Mycology, 9(2), 93–105.
Sousa, J., Silvério, S. C., Costa, A. M. A., & Rodrigues, L. R. (2022). Metagenomic approaches as
a tool to unravel promising biocatalysts from natural resources: Soil and water. Catalysts, 12,
385. Enzymes and Biocatalysis, 167.
Role of Bacterial Degradation in Lignocellulosic Biomass for Biofuel Production 315

Srivastava, N., Srivastava, M., Mishra, P. K., Ramteke, P. W., & Singh, R. L. (Eds.). (2019). New
and future developments in microbial biotechnology and bioengineering: From cellulose to
cellulase: Strategies to improve biofuel production. Elsevier.
Svetlitchnyi, V. A., Svetlichnaya, T. P., Falkenhan, D. A., Swinnen, S., Knopp, D., & Läufer,
A. (2022). Direct conversion of cellulose to L-lactic acid by a novel thermophilic
Caldicellulosiruptor strain. Biotechnology for Biofuels and Bioproducts, 15(1), 1–12.
Swathy, R., Rambabu, K., Banat, F., Ho, S. H., Chu, D. T., & Show, P. L. (2020). Production and
optimization of high grade cellulase from waste date seeds by Cellulomonas uda NCIM 2353
for biohydrogen production. International Journal of Hydrogen Energy, 45(42), 22260–22270.
Toussaint, M., Bontemps, C., Besserer, A., Hotel, L., Gérardin, P., & Leblond, P. (2016). Whole-­
cell biosensor of cellobiose and application to wood decay detection. Journal of Biotechnology,
239, 39–46.
Vaithanomsat, P., Trakunjae, C., Meelaksana, J., Boondaeng, A., Apiwatanapiwat, W., Janchai,
P., Boonyarit, J., & Chollakup, R. (2022). Improvement of pineapple leaf fiber quality by pec-
tinase produced from newly isolated Bacillus subtilis subsp. inaquosorum P4-1. Fibers and
Polymers, 23(2), 576–588.
Villota, E. M., Dai, Z., Lu, Y., & Yang, B. (2020). Enzymes for cellulosic biomass hydrolysis
and saccharification. In Green energy to sustainability: strategies for global industries
(pp. 283–326). Wiley.
Wu, D., Wei, Z., Mohamed, T. A., Zheng, G., Qu, F., Wang, F., Zhao, Y., & Song, C. (2022).
Lignocellulose biomass bioconversion during composting: Mechanism of action of lignocel-
lulase, pretreatment methods and future perspectives. Chemosphere, 286, 131635.
Yadav, K. K., Patil, P. B., Kumaraswamy, H. H., & Kashyap, B. K. (2020). Ligninolytic microbes
and their role in effluent management of pulp and paper industry. In Waste to energy: Prospects
and applications (pp. 309–350). Springer.
Yudden, N. K. M., Nordin, N. N., Harun, A., Rosli, N. H., Aziz, N. A., & Daud, S. (2019).
Comparative study of antioxidant activity of stem and leaves of Entada spiralis and their anti-
bacterial properties against Erwinia chrysanthemi. Science Letters, 13(2), 37–47.
Zeng, W., Liu, B., Wu, W., Zhang, S., Chen, Y., & Li, Z. (2021). Collaborative response of the
host and symbiotic lignocellulytic system to non-lethal toxic stress in Coptotermes formosanus
Skiraki. Insects, 12(6), 510.
Zuliani, L., Serpico, A., De Simone, M., Frison, N., & Fusco, S. (2021). Biorefinery gets hot:
Thermophilic enzymes and microorganisms for second-generation bioethanol production.
Processes, 9(9), 1583.
Zuo, S. S., Niu, D. Z., Ning, T. T., Zheng, M. L., Jiang, D., & Xu, C. C. (2018). Protein enrichment
of sweet potato beverage residues mixed with peanut shells by Aspergillus oryzae and Bacillus
subtilis using central composite design. Waste and Biomass Valorization, 9, 835–844.
Cultivating a Greener Tomorrow:
Sustainable Agriculture Strategies
for Minimizing Agricultural Waste

Dipti Bharti, Abhilekha Sharma, Meenakshi Sharma, Rahul Singh,


Amit Kumar, and Richa Saxena

1 Introduction

From 3.7 billion people in 1970 to 7.9 billion in 2021, the world’s population has
increased. By 2050 and 2100, respectively, it is expected to reach 9 and 11 billion
people (Koop & van Leeuwen, 2017). As a result, food security will provide a sig-
nificant issue in the future (Godfray et al., 2010). The production of livestock and
crops has increased significantly to meet the high demand of the growing popula-
tion, which has also increased the production of agricultural wastes (AWs) (Boserup,
1975; Tripathi et al., 2019). China, India, and Africa have all witnessed fast popula-
tion growth, economic expansion, and an increase in AW production capacities dur-
ing the past century (Wang et al., 2018).
India produces a lot of solid waste annually, with AW continuing to be the leader
with between 350 and 990 Mt/y. India is the second largest producer of agricultural
waste in the world after China, producing more than 130 million tonnes of paddy
straw annually, of which half is used as animal feed and the other half is discarded
(Singh & Sidhu, 2014). Additionally, the practice of burning rice bran (parali) in the

D. Bharti (*) · R. Singh


Department of Applied Science & Humanities, Darbhanga College of Engineering,
Darbhanga, Bihar, India
A. Sharma
Department of Chemistry, Noida International University, Greater Noida, Uttar Pradesh, India
M. Sharma
Department of Chemistry, IEC University, Baddi, Himachal Pradesh, India
A. Kumar
Department of Mechanical Engineering, Indian Institute of Technology, Ropar, Punjab, India
R. Saxena
Department of Biotechnology, Invertis University, Bareilly, Uttar Pradesh, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 317
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1_18
318 D. Bharti et al.

northwest region causes significant air pollution and poses a risk to the public’s
health (Shyamsundar et al., 2019). According to (Searchinger et al., 2008) and Kaab
et al. (2019), improper crop residue disposal produces greenhouse gases (GHGs)
such as carbon dioxide (CO2), nitrous oxide (N2O), and methane (CH4), which are
dangerous for both humans and the environment.
Animal waste (urine, excrement, wash water, leftover milk, and waste feed), bird
trash (spilled feed, feathers, droppings, and bedding material), slaughterhouse waste
(blood, hair, hides, flesh, bones, etc.), crop residues (leaf litter, seed pods, stalks,
stems, straws, husks, and weeds), and other waste products are all examples of
waste materials and agricultural waste (such as bagasse, molasses, and peels from
various fruits and vegetables, including orange and pota) (Seidavi et al., 2019;
Tripathi et al., 2019; Duque-Acevedo et al., 2020).
The high biomass produced by agriculture, one of the primary biological indus-
tries, can help the bioeconomy. According to Agamuthu (2009) and Bracco et al.
(2018), bioeconomic solutions based on AW management (AWM) can protect
farmer livelihood, chances for youth employment, and sustainability in agriculture
by preventing the careless or random burning of agricultural wastes and the under-
utilization of livestock manure. Most of the AWs listed above are simple to decom-
pose, and the end products not only provide essential nutrients for plants but also
encourage soil aeration and water retentivity by making the soil porous. In order to
contribute to a clean, safe, and sustainable agriculture, it will reduce GHG pollution
and the dependence on fossil fuels while also generating green markets and job
opportunities by turning agricultural wastes and by-products into valuable resources
(Mohanty et al., 2002; Scarlat et al., 2015, 2018). As a result, it is critical to reduce,
repurpose, and recycle agricultural waste in order to prevent pressures on soil, bio-
diversity, and global food security as well as to decouple environmental pressures
from economic growth (both resource and impact decoupling).
A significant portion of agricultural biomass is composed of lignocellulose,
which is made up of cellulose, hemicellulose, and lignin. Cellulose is a useful
resource for microbial metabolism because it makes up the majority of total bio-
mass (30–50%). Studies have shown that lignocellulosic biomass can be strategi-
cally managed and valorized to produce a number of goods with domestic and
commercial value. Compost, biocoal, charcoal, biobricks, biohydrogen, biometh-
ane, bioethanol, biobutanol, organic acids, and bioelectricity are a few examples.
Additionally, it has opened new doors for youth from farming communities around
the world who are looking for employment and will continue to do so (Sreekrishnan
et al., 2004; Sabiiti, 2011; Wahid & Nararya, 2015; Chi et al., 2014; Nigussie et al.,
2015; Koul et al., 2022). The public and farming communities need to be educated
more about the unintended advantages of biological and biotechnological manage-
ment of AWs, such as improved human health, less or no soil, air, or water pollution,
and alternative sources of income. By doing so, preconceived beliefs and scare tac-
tics based on ambiguous information would be eliminated (Westerman &
Bicudo, 2005).
This review is unique and different from past papers since it intends to increase
environmental awareness among the scientific community regarding the possibility
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 319

of reducing, reusing, and recycling AWs. It concentrates on the primary AW types


and sources and highlights the potential environmental issues that AWs may cause.
It promotes the use of agricultural wastes (AWs) as a vast resource library and out-
lines all practical strategies (traditional, biological, and biotechnological) for their
utilization and valorization into value-added products that could foster economic
growth, offer job opportunities to young people in farming communities, enrich
soil, produce abundant harvests without yield penalties, and ensure sustainable agri-
culture for the long term. It examines the advantages and disadvantages of the cur-
rent AW management systems and discusses the factors that influence the recycling
and utilization of AWs. It also gives policy suggestions for government-funded AW
management programs with a circular economy focus.

2 Unveiling Agricultural Waste: Exploring Types


and Sources

At many points along the supply chain, from production to consumption, agricul-
tural waste is produced. An overview of several forms of agricultural waste and their
sources is given below:
1. Crop Residues: After harvesting crops, residues such as stems, leaves, and
husks remain in the field. These residues can accumulate and contribute to
waste if not managed properly.
2. Food Waste: Food waste is produced at different points in the supply chain,
including during post-harvest handling, processing, distribution, and consump-
tion. Unsold produce, expired products, and food scraps are common forms of
food waste.
3. Packaging Materials: Packaging materials, including plastics, cardboard, and
containers, are used to protect and transport agricultural products. However,
these materials can become waste if not recycled or disposed of properly.
4. Agrochemical Containers: Containers used for pesticides, fertilizers, and other
agrochemicals can generate waste when they are discarded after use.
5. Animal Manure: Livestock farming produces substantial amounts of animal
manure, which can be a valuable source of nutrients but also poses waste man-
agement challenges.
6. Processing By-Products: Agricultural processing generates by-products, such
as peels, shells, and husks, that are not used in the final product. These by-­
products contribute to waste.
7. Unsold Produce: Produce that does not meet market standards or remains
unsold due to various reasons contributes to waste.
8. Expired Inputs: Agricultural inputs such as fertilizers, pesticides, and seeds can
become waste when they expire or are no longer usable.
320 D. Bharti et al.

9. Non-Composted Biomass: Organic waste from agricultural activities, such as


non-composted plant material and crop residues, can become waste if not man-
aged through proper composting.
10. Weeds and Invasive Species: During weed management, removed weeds and
invasive species become organic waste that requires proper disposal (Sheikh
et al., 2022; Parfitt et al., 2010).

3 Nurturing Tomorrow’s Harvest: Crafting Policy


Considerations for Sustainable Agriculture

Recently, the phrase “sustainability” has gained popularity in relation to agricultural


initiatives in terms of teaching, extension, research, and public policy. There must
be some degree of consensus over the definition of the phrase “sustainable agricul-
ture” for there to be efficient communication among those who teach, study, and
practice agriculture. The definition must be broad enough to cover the large range of
agricultural settings to which it will be applied, while also being precise enough to
give standards by which various systems’ viability may be assessed. The definition
should be “ends oriented” rather than “means oriented” to provide the maximum
possible combination of utility and flexibility.
Though concept has grown in prominence in politics and academia, sustainabil-
ity can mean different things to different individuals.
People who teach, study, practice, regulate, or support agriculture must have a
basic understanding of what is meant by the term “sustainable agriculture.”
Otherwise, government and business initiatives to promote farmer transition and
sustainable agricultural research can just wind up adopting outmoded standards
under a new name.
The types of agricultural systems and practices that should be included in courses
on the subject will also be difficult for professors and students to agree on in the
absence of a defined definition. The definition must be broad enough to include the
variety of agricultural conditions in which it will be used, while also being detailed
enough to provide standards by which a proposed system’s sustainability may be
assessed.
As a result, the concept of sustainable agriculture seems to be founded on three
main areas of concern: (i) economic concerns about economic justice, the survival
of owner-operated farms, and the industry’s long-term profitability; (ii) environ-
mental concerns about the negative effects of agriculture on land, water, and wild-
life resources; and (iii) public welfare concerns about food safety and human
exposure to toxic chemicals. The sustainable agricultural plan differs from more
conventional techniques in terms of integration level.
In a sustainable farm, all system components are integrated in the horizontal,
vertical, and temporal dimensions. The effects of fieldscale systems on farm, water-
shed, regional, and national levels must be assessed in the horizontal dimension, and
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 321

vice versa. (For example, the effects of national policy on local communities and
farms must be assessed.) It is probably impossible to describe sustainable agricul-
ture in terms of particular sets of practices because agriculture comprises processes
that are site- and situation-specific. Furthermore, as sustainability depends on social,
political, and economic variables, it is impossible to define sustainable agriculture
without taking these factors into consideration (Altieri, 2002). Agriculture is
dynamic by both biological and social nature.
A sustainable agriculture is one that, over the long term, improves environmental quality
and the resource base on which agriculture depends; provides for basic human food and
fiber needs; is economically viable; and enhances the quality of life for farmers and society
as a whole. (Neher, 2018)

If an agricultural program, policy, or practice:


1. In a sustainable farm, all system components are integrated in the horizontal,
vertical, and temporal dimensions. The effects of fieldscale systems on farm,
watershed, regional, and national levels must be assessed in the horizontal
dimension, and vice versa.
2. Enhances rather than degrades the managed agricultural ecosystem’s long-term
production, diversity, and integrity, as well as the nearby natural ecosystems.
3. Improves rather than jeopardizes the aesthetic satisfaction, safety, and health of
both agricultural producers and consumers (Fig. 1).

4 Seeds of Change: Guiding Principles for a Flourishing


Future in Sustainable Agriculture

A promising solution to the problems caused by agricultural waste is to practice


sustainable agriculture. By maximizing resource usage, reducing waste creation,
and promoting environmentally friendly practices, these approaches support waste
reduction objectives.
Organic farming: Organic farming aims to reduce the use of synthetic inputs
while promoting soil health and biodiversity. Waste is decreased by the following:

Composting: Closing nutrient loops and reducing waste by turning organic waste
(such as crop residues and manure) into nutrient-rich compost.
Reduced Chemical Use: Organic farming decreases the possibility of chemical resi-
dues and contamination by staying away from synthetic pesticides and fertilizers
(Reganold & Wachter, 2016).
IPM, or integrated pest management: It is the practice of managing pests while
reducing their negative effects on the environment. Waste is decreased by the
following:
322 D. Bharti et al.

Environmental
Sustainability
Environmental Safety
Land Degradation
Air & Water Degradation
Emission of CO2

Social Sustainability Economic Sustainability


Resource Optimization Crop Production Cost
Quality of Life Crop Productivity
Migrating From Below Poverty Line (BPL) Crop Production Rate
Food Self Sufficiency Employment & Self Reliance
Know – How & Skill Farm Economy

Fig. 1 Strategies of sustainable agricultural practices

Targeted Approaches: IPM uses targeted treatments to control pests, which elimi-
nates the need for using broad-spectrum chemicals.
Monitoring and Prevention: Regular monitoring makes it possible to spot problems
before they become serious, hence reducing crop losses and waste.
Agroecology: Agroecological methods place an emphasis on how agriculture and
ecosystems are interconnected. They lessen waste through the following:

Diverse Cropping Systems: Polyculture and crop rotation lower the likelihood that
crops will fail due to pests or diseases, reducing waste.
Enhancing natural processes such as pollination and nutrient cycling cuts down
on the demand for outside inputs and waste-producing activities (Gurr et al., 2016).
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 323

5 Turning Trash into Treasure: Harnessing Resources


from Waste for a Sustainable Tomorrow

Agricultural waste may be converted into useful resources, providing a long-term


waste management solution and opening up new business prospects. Agricultural
waste can be transformed into goods such as bioenergy, bioplastics, and animal feed
using a variety of waste-to-resource techniques.
Bioenergy Production: Techniques such as anaerobic digestion and biomass gas-
ification can turn agricultural waste into bioenergy. In the absence of oxygen, anaer-
obic digestion decomposes organic waste, creating biogas containing methane.
Then, this biogas can be used to generate energy and heat. Waste biomass is ther-
mally converted into synthetic gas (syngas) by the process of biomass gasification,
which can then be utilized to produce electricity or biofuels (Parawira, 2009).

Bioplastics and Biomaterials: Agricultural waste, which is abundant in organic


components such as cellulose and lignin, can be used as a feedstock for the produc-
tion of bioplastics and biomaterials. These environmentally friendly replacements
for conventional plastics can lessen our dependency on fossil fuels and cut down on
plastic waste. Waste-derived feedstocks can be converted into biodegradable plas-
tics and other bio-based materials through processes including fermentation and
polymerization.

Animal Feed and Nutrient Recovery: Certain agricultural wastes can be pro-
cessed and used as animal feed. This is known as “nutrient recovery.” For instance,
agricultural leftovers and by-products of food processing can be transformed into
nutrient-dense cattle feed, lowering the demand for conventional feed sources. In
addition, recovering nutrients from waste, such phosphate and nitrogen, might be
useful for fertilizing crops (Cordell et al., 2009).

6 Case Studies for Waste-to-Resource Projects That Worked

Anaerobic digestion for the production of biogas: The Indian state of Kerala estab-
lished a biogas program that produced biogas for cooking and electricity generation
from organic waste and agricultural waste. This program enhanced rural popula-
tions’ access to energy while addressing trash management.
Waste-Derived Bioplastics: The Mater-Bi line of bioplastics was developed by
the Italian bioplastics business Novamont. They lessen the carbon footprint of plas-
tics and support the concepts of the circular economy by using agricultural by-­
products as feedstock, such as maize starch and vegetable oils (Chiellini et al., 2003).
324 D. Bharti et al.

7 Revolutionizing Waste Reduction: Exploring Innovations


and Technologies for a Sustainable Future

Technology and innovation are essential in the modern agricultural environment for
minimizing waste and improving farming methods. The Internet of Things (IoT),
digital tools, and sensors are being used to increase resource efficiency, forecast
waste production, and stop losses.
Digital Tools, Sensors, and the Internet of Things: Digital platforms and sensors
allow for real-time data gathering and analysis, which results in well-informed deci-
sions. IoT combines sensors and gadgets, enabling remote monitoring and manage-
ment of numerous farming businesses’ processes. This technique helps with
resource allocation optimization, lowering input overuse, and boosting overall
effectiveness (Turner et al., 2015).

Emerging Technologies for Waste Monitoring: Non-invasive approaches to


evaluate crop health and spot possible problems are provided by advanced imaging
techniques including hyperspectral and multispectral imaging. High-resolution
images from drones with sensors allows for precise field monitoring and the early
identification of insect outbreaks or nutrient deficits (Odebiri et al., 2020).

Waste Prediction and Prevention: Along the agricultural supply chain, data-­
driven models and machine learning algorithms are being created. These models
forecast variables such as weather conditions or changes in market demand by ana-
lyzing historical data and current information.
By incorporating these technologies, farmers can make more informed deci-
sions, optimize resource use, and take proactive measures to prevent waste, ulti-
mately contributing to more sustainable and efficient agricultural practices.

8 Sustainable Agriculture Versus Sustainability Driven


by Agriculture

Sustainable Agriculture: A holistic approach that attempts to address the present


and future demands of agriculture while taking into account economic, social, and
environmental factors is referred to as sustainable agriculture. It includes methods
for enhancing long-term productivity without harming the environment, preserving
the health of the ecosystem, and enhancing the welfare of farming communities.
Crop rotation, integrated pest management, and soil conservation are examples of
sustainable agriculture practices that aim to reduce environmental harm while main-
taining robust and resilient agricultural systems (Pretty, 2008).
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 325

Agriculture-Driven Sustainability: This phrase describes how agricultural prac-


tices have a larger impact on overall sustainability objectives. While sustainability
driven by agriculture accepts that agriculture can have a substantial impact on more
general sustainability goals, sustainable agriculture focuses on the practices and
systems within the agricultural sector (Erb et al., 2016). This covers topics including
reducing global warming, conserving biodiversity, managing water resources, and
rural development. Agriculture has the potential to benefit efforts to achieve global
sustainability because it uses a lot of resources and takes up a lot of area (Smith
et al., 2014).

9 Balancing Choices: A Comparative Exploration

The distinction between sustainable agriculture and sustainability driven by agricul-


ture lies in their scope and focus. Sustainable agriculture primarily addresses on-­
farm practices and systems, seeking to optimize agricultural productivity while
minimizing negative impacts. Contrarily, sustainability fueled by agriculture
acknowledges that the agricultural sector’s actions can have significant effects on
global sustainability, encompassing aspects of the environment, society, and the
economy.
While sustainable agriculture is a crucial component of sustainability driven by
agriculture, the latter emphasizes the broader implications of agricultural practices,
influencing policies, international agreements, and multi-sectoral collaborations to
achieve comprehensive sustainability goals.
In essence, sustainable agriculture is a key building block that contributes to
sustainability driven by agriculture, which, in turn, addresses the overarching goal
of creating a more sustainable and equitable global future through agricultural prac-
tices and their interconnected effects.

10 Unlocking Nutrient Potential: Exploring Widely Adopted


Processes for Nutrient Recovery from Waste

Nutrient recovery processes from waste play a crucial role in minimizing environ-
mental pollution, conserving resources, and promoting circular economy principles.
Here, we delve into several commonly used nutrient recovery processes, highlight-
ing their significance and providing references for further exploration.
In the biological process of anaerobic digestion, organic material is broken down
by bacteria without the use of oxygen, producing biogas (mostly methane) and a
residue material called digestate that is rich in nutrients. Nitrogen and phosphorus
in the digestate are significant sources of nutrients that can be used as fertilizers
(Parawira, 2009).
326 D. Bharti et al.

Struvite Precipitation: Struvite, a compound containing magnesium, ammonium,


and phosphate, can be recovered from wastewater or waste streams through con-
trolled precipitation. This struvite can be used as a slow-release fertilizer to supply
crops with vital nutrients.
Algal Cultivation: Nutrient-rich waste streams, such as wastewater and agricul-
tural runoff, can be used to grow algae. An excellent source of nutrients, notably
nitrogen and phosphorus, can be found in the collected algal biomass. Algal bio-
mass can also be utilized for the manufacture of high-value goods or biofuels
(Brennan & Owende, 2010).
Composting: Composting is the regulated natural degradation of organic waste
products. The compost that is produced is abundant in organic material, nitrogen,
and phosphorus. Compost is a soil supplement that improves soil structure and fer-
tility (Sáez et al., 2021).
Insect Farming (e.g., feeding black army flies organic waste): Insect farming
entails feeding insects, like black soldier flies, organic waste. The waste is trans-
formed by the insects into biomass that is protein-rich and useful as animal feed and
a source of nutrition (Diener et al., 2009) (Table 1).

11 Barriers to Adoption and Solutions

Socioeconomic Challenges Faced by Farmers in Adopting Sustainable


Practices
When switching to sustainable farming practices, farmers may face a variety of
socioeconomic obstacles. These difficulties may include restricted access to finan-
cial resources, a lack of technical expertise, and uncertainty over market demand for
products made with sustainability in mind (Dorward et al., 2003).
Strategies to Overcome Barriers Through Education, Training, and Support
Education, training, and support systems are essential for tackling these prob-
lems. Programs for farmer education, extension services, and training seminars
can give farmers the information and abilities they need to successfully apply
sustainable practices. Economic obstacles can also be reduced through financial
incentives, such as grants or subsidies for adopting sustainable technologies
(Ezekannagha, 2020).

12 Success Stories and Case Studies

Examples of Regions, Farms, or Organizations That Have Effectively


Reduced Agricultural Waste
Successful waste reduction techniques in agriculture have been demonstrated by a
number of areas, farms, and organizations. The Netherlands, for instance, has built
circular agricultural systems that maximize resource usage and reduce waste. By
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 327

Table 1 Overview of commonly used nutrient recovery processes from waste


Nutrient recovery Recovered
process Description Waste source nutrients Reference
The anaerobic Biogas and digestate are Organic Biogas (methane), Bouallagui
process produced by the waste, digestate rich in et al. (2005)
microbial decomposition wastewater nitrogen and
of organic waste in the sludge phosphorus
absence of oxygen
Composting Natural decomposition Organic Compost rich in Diaz et al.
of organic waste by waste, organic matter, (2005)
microorganisms, agricultural nitrogen,
yielding nutrient-rich residues phosphorus, and
compost micronutrients
Algal cultivation Growing algae on Wastewater, Algal biomass rich Converti
nutrient-rich waste agricultural in nitrogen, et al. (2009)
streams, harvesting runoff phosphorus, and
algae biomass for sometimes lipids
nutrient recovery
Struvite Precipitation of struvite Wastewater, Struvite crystals Siciliano
precipitation (magnesium ammonium digestate rich in nitrogen and and Rosa
phosphate) from phosphorus (2014)
wastewater, forming
slow-release fertilizer
Hydrothermal Conversion of organic Organic Biocrude oil Conti et al.
liquefaction waste into biocrude oil waste, sewage containing carbon (2020)
through high sludge and nutrients
temperature and
pressure
Insect farming Feeding organic waste to Organic Insect biomass rich Diener et al.
(e.g., Black insects, harvesting insect waste, food in protein, fats, and (2009)
soldier fly) biomass for nutrient scraps other nutrients
extraction
Nutrient Using membranes to Wastewater, Concentrated Kumar and
extraction separate and concentrate agricultural nutrient solutions Pal (2015)
membranes nutrients from liquid runoff (nitrogen,
waste streams phosphorus, etc.)
Acid hydrolysis Treatment of organic Organic Nutrient-rich liquid Benatti and
waste with acids to waste, food solution Polizeli
break down complex processing (2023)
molecules and release residues
nutrients
Bioleaching Using microorganisms Electronic Extracted metals Bosecker
to extract metals and waste, mine and sometimes (1997)
nutrients from waste tailings associated nutrients
materials
Electrochemical Applying electrical Various Recovered nutrient Zhang et al.
recovery currents to separate and wastewater ions (nitrogen, (2023)
recover nutrients from sources phosphorus, etc.)
waste streams
328 D. Bharti et al.

reusing garbage for composting or energy production, several farms have adopted
zero-waste policies. Food waste can be decreased because to programs created by
organizations such as “Food Forward” in the United States that rescue extra produce.
Lessons Learned and Transferrable Practices: A number of lessons and transfer-
rable practices can be learned from these success stories. Integrated strategies that
take into account economic, environmental, and social aspects frequently produce
superior outcomes. Governmental organizations, academic institutions, and regional
communities working together can successfully bring about change. The demand
for goods made responsibly can also be increased by educating customers about
sustainable consumption habits (Moshontz et al., 2018).

13 Horizons Unveiled: Shaping the Future and Charting


Research Avenues

Potential Advancements in Waste Reduction Through Ongoing Research


1. Advanced Sensor Technologies: Continued development of sensors and remote
sensing technologies can enable real-time monitoring of crops, soil conditions,
and waste streams. This will enhance precision agriculture practices and aid in
early detection of potential waste-related issues (Moshou et al., 2005).
2. Big Data and Artificial Intelligence (AI): The integration of big data analytics
and AI can lead to more accurate predictions of waste generation, pest outbreaks,
and resource optimization. Machine learning algorithms can provide insights
into complex relationships within agricultural systems, supporting waste reduc-
tion strategies (Zhang et al., 2020).
3. Circular Economy Practices: Further exploration of circular economy principles
in agriculture, such as nutrient cycling, resource reuse, and waste valorization,
can lead to innovative solutions for waste reduction and resource efficiency.

14 Unveiling Significance: Illuminating the Importance


of Our Endeavors

The importance of a review article on sustainable agricultural methods for cutting


down on agricultural waste lies in its capacity to thoroughly summarize and synthe-
size existing research, offering beneficial insights and direction for diverse stake-
holders. The following main ideas underline the importance of this chapter:
Knowledge Consolidation: For the purpose of minimizing agricultural waste, a
review article combines several studies, approaches, and conclusions in one place.
It offers readers a central location to acquire information that is condensed and cur-
rent and covers a wide range of disciplines. Policymakers, agricultural extension
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 329

agents, and government organizations can use the knowledge gained from this
­analysis to create knowledgeable policies and strategies for supporting sustainable
practices (Blay-Palmer et al., 2016).
Farmers need to be informed since they are at the cutting edge of agricultural
techniques. With the help of this assessment, they will have a thorough understand-
ing of sustainable practices that can increase productivity, decrease waste, and make
operations more resilient.
Promoting Sustainable Research: This review can help researchers find gaps in
the field’s body of knowledge. It may stimulate additional investigation into novel
procedures, tools, and strategies for efficiently dealing with agricultural waste.
Increasing food production and availability is one way that sustainable farming
methods address the issue of food security by reducing waste. Because of how cru-
cial it is to use resources efficiently as the world’s population grows, this is extremely
crucial.
Environmental Conservation: The assessment places a strong emphasis on meth-
ods for reducing environmental harm, such as greenhouse gas emissions, water pol-
lution, and soil erosion. The ecosystems and natural resources are preserved as a
result of these actions.

Business and Economic Benefits: Using resources more effectively, spending less
on inputs, and improving soil health can all result in cost savings when sustainable
practices are adopted. These economic advantages for farmers and agribusinesses
might be highlighted by this review.

Promoting Innovation: The review may reveal new developments and inventions
in the fields of waste reduction and sustainable agriculture. Researchers, politicians,
and practitioners may work together as a result to create and put into practice
cutting-­edge solutions. The evaluation can be used as a teaching resource by aca-
demic institutions to increase student, researcher, and educator knowledge of and
comprehension of sustainable agriculture methods (Mariyono, 2020).

Contribution to the Global Agenda: The Sustainable Development Goals (SDGs)


give sustainable agriculture top importance. By offering information on minimizing
agricultural waste, this assessment aids in the worldwide drive to achieve SDG 2
(Zero Hunger) and SDG 12 (Responsible Consumption and Production) (Wood
et al., 2018).
A review article on sustainable agricultural methods for minimizing agricultural
waste is extremely important for developing knowledge, assisting in making deci-
sions, and advocating methods that are in line with environmental preservation and
sustainable development objectives.
330 D. Bharti et al.

15 Conclusion: Paving the Path to a Greener Tomorrow

The importance of adopting sustainable strategies to reduce agricultural waste can-


not be emphasized as we stand at the nexus of environmental stewardship and agri-
cultural growth. The journey we took, which is described in this review’s pages,
reveals a profound awareness of the difficulties and opportunities that lie ahead. We
have looked into a world where every grain stalk, leaf, and drop of water contains
the potential for transformation through the lenses of innovation, research, and
collaboration.
This review has knitted together strands of expertise, experience, and vision,
from the vast fields of crop rotation to the intricate networks of precision farming,
from the symphony of agroforestry to the microscopic world of nutrient recovery. It
has shed light on the interactions between societal equality, environmental resil-
ience, and economic growth within the complex web of sustainable agriculture.
Our research has extended beyond the confines of fields and laboratories, touch-
ing on social issues and political spheres. We have opened the road for a paradigm
change by revealing the fundamental worth of agricultural waste—one that supports
circular economies, uncouples expansion from environmental damage, and fosters
wealth alongside the health of the world. We are currently in a position where we
are trusted with a worthy cause. Aspiring not only for innovation but also for a
future worthy of future generations, the objective is to cultivate not only crops but
also landscapes, economies, and livelihoods. This study urges all stakeholders—
farmers, policymakers, researchers, and consumers—to embrace a journey that car-
ries within it the promise of a greener tomorrow. It is a tribute to the unrelenting
pursuit of knowledge and the tenacity of the human spirit. With the knowledge
gained here, let us move forward and plant the seeds of change in a future where
every harvest has the capacity to produce solutions and where every trash may be
converted into a resource. Let us plant the seeds of collaboration, nurture the growth
of innovation, and reap the rewards of sustainability. Together, we can pave a way
that feeds not only our plates but also the very fabric of our planet. Together, let us
create a world where agricultural waste serves as a beacon, pointing the way to a
prosperous and peaceful existence rather than being a burden.

References

Agamuthu, P. (2009). Challenges and opportunities in agro-waste management: An Asian per-


spective. In Inaugural meeting of first regional 3R forum in Asia (Vol. 11, 12). University of
Malaysia.
Altieri, M. A. (2002). Agroecology: The science of natural resource management for poor farmers
in marginal environments. Agriculture, Ecosystems & Environment, 93(1–3), 1–24.
Benatti, A. L. T., & Polizeli, M. D. L. T. D. M. (2023). Lignocellulolytic biocatalysts: The
main players involved in multiple biotechnological processes for biomass valorization.
Microorganisms, 11(1), 162.
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 331

Blay-Palmer, A., Sonnino, R., & Custot, J. (2016). A food politics of the possible? Growing sustain-
able food systems through networks of knowledge. Agriculture and Human Values, 33, 27–43.
Bosecker, K. (1997). Bioleaching: Metal solubilization by microorganisms. FEMS Microbiology
Reviews, 20(3–4), 591–604.
Boserup, E. (1975). The impact of population growth on agricultural output. The Quarterly Journal
of Economics, 89(2), 257–270.
Bouallagui, H., Touhami, Y., Cheikh, R. B., & Hamdi, M. (2005). Bioreactor performance in
anaerobic digestion of fruit and vegetable wastes. Process Biochemistry, 40(3–4), 989–995.
Bracco, S., Calicioglu, O., Gomez San Juan, M., & Flammini, A. (2018). Assessing the contribu-
tion of bioeconomy to the total economy: A review of national frameworks. Sustainability,
10(6), 1698.
Brennan, L., & Owende, P. (2010). Biofuels from microalgae—A review of technologies for pro-
duction, processing, and extractions of biofuels and co-products. Renewable and Sustainable
Energy Reviews, 14(2), 557–577.
Chi, X., Wang, M. Y., & Reuter, M. A. (2014). E-waste collection channels and household recy-
cling behaviors in Taizhou of China. Journal of Cleaner Production, 80, 87–95.
Chiellini, E., Corti, A., & Swift, G. (2003). Biodegradation of thermally-oxidized, fragmented
low-density polyethylenes. Polymer Degradation and Stability, 81(2), 341–351.
Conti, F., Toor, S. S., Pedersen, T. H., Seehar, T. H., Nielsen, A. H., & Rosendahl, L. A. (2020).
Valorization of animal and human wastes through hydrothermal liquefaction for biocrude
production and simultaneous recovery of nutrients. Energy Conversion and Management,
216, 112925.
Converti, A., Casazza, A. A., Ortiz, E. Y., Perego, P., & Del Borghi, M. (2009). Effect of tempera-
ture and nitrogen concentration on the growth and lipid content of Nannochloropsis oculata and
Chlorella vulgaris for biodiesel production. Chemical Engineering and Processing: Process
Intensification, 48(6), 1146–1151.
Cordell, D., Drangert, J. O., & White, S. (2009). The story of phosphorus: Global food security and
food for thought. Global Environmental Change, 19(2), 292–305.
Diaz, L. F., Savage, G. M., & Eggerth, L. L. (2005). Alternatives for the treatment and disposal of
healthcare wastes in developing countries. Waste Management, 25(6), 626–637.
Diener, S., Zurbrügg, C., & Tockner, K. (2009). Conversion of organic material by black soldier
fly larvae: Establishing optimal feeding rates. Waste Management & Research, 27(6), 603–610.
Dorward, A., Poole, N., Morrison, J., Kydd, J., & Urey, I. (2003). Markets, institutions and tech-
nology: Missing links in livelihoods analysis. Development Policy Review, 21(3), 319–332.
Duque-Acevedo, M., Belmonte-Ureña, L. J., Cortés-García, F. J., & Camacho-Ferre, F. (2020).
Agricultural waste: Review of the evolution, approaches and perspectives on alternative uses.
Global Ecology and Conservation, 22, e00902.
Erb, K. H., Lauk, C., Kastner, T., Mayer, A., Theurl, M. C., & Haberl, H. (2016). Exploring the
biophysical option space for feeding the world without deforestation. Nature Communications,
7(1), 11382.
Ezekannagha, E. (2020). Assessing the climatic suitability of Bambara groundnut as an underuti-
lised crop to future climate projections in Sikasso and Ségou, Mali (Master’s thesis, Faculty
of Science).
Godfray, H. C. J., Beddington, J. R., Crute, I. R., Haddad, L., Lawrence, D., Muir, J. F., Pretty, J.,
Robinson, S., Thomas, S. M., & Toulmin, C. (2010). Food security: The challenge of feeding 9
billion people. Science, 327(5967), 812–818.
Gurr, G. M., Lu, Z., Zheng, X., Xu, H., Zhu, P., Chen, G., Yao, X., Cheng, J., Zhu, Z., Catindig,
J. L., & Villareal, S. (2016). Multi-country evidence that crop diversification promotes ecologi-
cal intensification of agriculture. Nature Plants, 2(3), 1–4.
Kaab, A., Sharifi, M., Mobli, H., Nabavi-Pelesaraei, A., & Chau, K. W. (2019). Combined life
cycle assessment and artificial intelligence for prediction of output energy and environmental
impacts of sugarcane production. Science of the Total Environment, 664, 1005–1019.
332 D. Bharti et al.

Koop, S. H., & van Leeuwen, C. J. (2017). The challenges of water, waste and climate change in
cities. Environment, Development and Sustainability, 19(2), 385–418.
Koul, B., Yakoob, M., & Shah, M. P. (2022). Agricultural waste management strategies for envi-
ronmental sustainability. Environmental Research, 206, 112285.
Kumar, R., & Pal, P. (2015). Assessing the feasibility of N and P recovery by struvite precipitation
from nutrient-rich wastewater: A review. Environmental Science and Pollution Research, 22,
17453–17464.
Mariyono, J. (2020). Improvement of economic and sustainability performance of agribusiness
management using ecological technologies in Indonesia. International Journal of Productivity
and Performance Management, 69(5), 989–1008.
Mohanty, A. K., Misra, M., & Drzal, L. T. (2002). Sustainable bio-composites from renewable
resources: Opportunities and challenges in the green materials world. Journal of Polymers and
the Environment, 10, 19–26.
Moshontz, H., Campbell, L., Ebersole, C. R., IJzerman, H., Urry, H. L., Forscher, P. S., Grahe,
J. E., McCarthy, R. J., Musser, E. D., Antfolk, J., & Castille, C. M. (2018). The Psychological
Science Accelerator: Advancing psychology through a distributed collaborative network.
Advances in Methods and Practices in Psychological Science, 1(4), 501–515.
Moshou, D., Bravo, C., Oberti, R., West, J., Bodria, L., McCartney, A., & Ramon, H. (2005). Plant
disease detection based on data fusion of hyper-spectral and multi-spectral fluorescence imag-
ing using Kohonen maps. Real-Time Imaging, 11(2), 75–83.
Neher, D. (2018). Ecological sustainability in agricultural systems: Definition and measurement. In
Integrating sustainable agriculture, ecology, and environmental policy (pp. 51–61). Routledge.
Nigussie, A., Kuyper, T. W., & de Neergaard, A. (2015). Agricultural waste utilisation strategies
and demand for urban waste compost: Evidence from smallholder farmers in Ethiopia. Waste
Management, 44, 82–93.
Odebiri, O., Mutanga, O., Odindi, J., Peerbhay, K., & Dovey, S. (2020). Predicting soil organic
carbon stocks under commercial forest plantations in KwaZulu-Natal province, South Africa
using remotely sensed data. GIScience & Remote Sensing, 57(4), 450–463.
Parawira, W. (2009). Biogas technology in sub-Saharan Africa: Status, prospects and constraints.
Reviews in Environmental Science and Bio/Technology, 8, 187–200.
Parfitt, J., Barthel, M., & Macnaughton, S. (2010). Food waste within food supply chains:
Quantification and potential for change to 2050. Philosophical Transactions of the Royal
Society B: Biological Sciences, 365(1554), 3065–3081.
Pretty, J. (2008). Agricultural sustainability: Concepts, principles and evidence. Philosophical
Transactions of the Royal Society B: Biological Sciences, 363(1491), 447–465.
Reganold, J. P., & Wachter, J. M. (2016). Organic agriculture in the twenty-first century. Nature
Plants, 2(2), 1–8.
Sabiiti, E. N. (2011). Utilising agricultural waste to enhance food security and conserve the envi-
ronment. African Journal of Food, Agriculture, Nutrition and Development, 11(6), 1–9.
Sáez, J. A., Pérez-Murcia, M. D., Vico, A., Martínez-Gallardo, M. R., Andreu-Rodríguez, F. J.,
López, M. J., Bustamante, M. A., Sanchez-Hernandez, J. C., Moreno, J., & Moral, R. (2021).
Olive mill wastewater-evaporation ponds long term stored: Integrated assessment of in situ
bioremediation strategies based on composting and vermicomposting. Journal of Hazardous
Materials, 402, 123481.
Scarlat, N., Dallemand, J. F., & Fahl, F. (2018). Biogas: Developments and perspectives in Europe.
Renewable Energy, 129, 457–472.
Scarlat, N., Dallemand, J. F., Monforti-Ferrario, F., Banja, M., & Motola, V. (2015). Renewable
energy policy framework and bioenergy contribution in the European Union–An overview from
National Renewable Energy Action Plans and Progress Reports. Renewable and Sustainable
Energy Reviews, 51, 969–985.
Searchinger, T., Heimlich, R., Houghton, R. A., Dong, F., Elobeid, A., Fabiosa, J., Tokgoz, S.,
Hayes, D., & Yu, T. H. (2008). Use of US croplands for biofuels increases greenhouse gases
through emissions from land-use change. Science, 319(5867), 1238–1240.
Cultivating a Greener Tomorrow: Sustainable Agriculture Strategies for Minimizing… 333

Seidavi, A. R., Zaker-Esteghamati, H., & Scanes, C. G. (2019). Present and potential impacts
of waste from poultry production on the environment. World’s Poultry Science Journal,
75(1), 29–42.
Sheikh, M. R., Ali, N. A., & Aslam, A. (2022). Food wastage footprint, food security, environment
and economic growth nexus in developing countries. InTech.
Shyamsundar, P., Springer, N. P., Tallis, H., Polasky, S., Jat, M. L., Sidhu, H. S., Krishnapriya,
P. P., Skiba, N., Ginn, W., Ahuja, V., & Cummins, J. (2019). Fields on fire: Alternatives to crop
residue burning in India. Science, 365(6453), 536–538.
Siciliano, A., & Rosa, S. D. (2014). Recovery of ammonia in digestates of calf manure through a
struvite precipitation process using unconventional reagents. Environmental Technology, 35(7),
841–850.
Singh, Y., & Sidhu, H. S. (2014). Management of cereal crop residues for sustainable rice-wheat
production system in the Indo-Gangetic plains of India. Proceedings of the Indian National
Science Academy, 80(1), 95–114.
Smith, P., Bustamante, M., Ahammad, H., Clark, H., Dong, H., Elsiddig, E. A., Haberl, H.,
Harper, R., House, J., Jafari, M., & Masera, O. (2014). Agriculture, forestry and other land use
(AFOLU). In Climate change 2014: Mitigation of climate change (Contribution of Working
Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change)
(pp. 811–922). Cambridge University Press.
Sreekrishnan, T. R., Kohli, S., & Rana, V. (2004). Enhancement of biogas production from solid
substrates using different techniques––A review. Bioresource Technology, 95(1), 1–10.
Tripathi, N., Hills, C. D., Singh, R. S., & Atkinson, C. J. (2019). Biomass waste utilisation in low-­
carbon products: Harnessing a major potential resource. npj Climate and Atmospheric Science,
2(1), 35.
Turner, W., Rondinini, C., Pettorelli, N., Mora, B., Leidner, A. K., Szantoi, Z., Buchanan, G., Dech,
S., Dwyer, J., Herold, M., & Koh, L. P. (2015). Free and open-access satellite data are key to
biodiversity conservation. Biological Conservation, 182, 173–176.
Wahid, A., & Nararya, R. N. (2015). Optimasi Pengendalian Unit Gasifikasi dan Char Combustor
pada Pabrik Biohidrogen dari Biomassa Menggunakan Reidentifikasi Model Predictive
Control. In Prosiding Seminar Nasional Teknik Kimia Indonesia V (pp. 12–13). Universitas
Gadjah Mada.
Wang, F., Zhao, H., Xiang, H., Wu, L., Men, X., Qi, C., Chen, G., Zhang, H., Wang, Y., & Xian,
M. (2018). Species diversity and functional prediction of surface bacterial communities on
aging flue-cured tobaccos. Current Microbiology, 75, 1306–1315.
Westerman, P. W., & Bicudo, J. R. (2005). Management considerations for organic waste use in
agriculture. Bioresource Technology, 96(2), 215–221.
Wood, S. L., Jones, S. K., Johnson, J. A., Brauman, K. A., Chaplin-Kramer, R., Fremier, A.,
Girvetz, E., Gordon, L. J., Kappel, C. V., Mandle, L., & Mulligan, M. (2018). Distilling the role
of ecosystem services in the Sustainable Development Goals. Ecosystem Services, 29, 70–82.
Zhang, P., Li, Z., Ghardallou, W., Xin, Y., & Cao, J. (2023). Nexus of institutional quality and
technological innovation on renewable energy development: Moderating role of green finance.
Renewable Energy, 214, 233–241.
Zhang, X., He, L., Zhang, J., Whiting, M. D., Karkee, M., & Zhang, Q. (2020). Determination of
key canopy parameters for mass mechanical apple harvesting using supervised machine learn-
ing and principal component analysis (PCA). Biosystems Engineering, 193, 247–263.
Index

A Biomass, 1, 36, 51, 69, 99, 115, 133, 154, 171,


Agricultural solid waste, 57, 317 189, 208, 224, 239, 269, 303, 318
Agriculture, 25, 62, 70, 149, 175, 192, 209, Biomass energy, 1–14, 37, 184
223, 239, 263, 269, 296, 318 Biomass management, 69–94, 189–202,
Agro-wastes, 125, 133, 134, 138, 207–219 223–232, 239–250
Application of enzymes, 81, 189–202, 212, Biomass valorization, 99–111, 134
225, 232 Biomass wastes, 1–4, 10, 55, 69–94, 115–129,
171–184, 189–202,
223–232, 239–250
B Biowaste, 51–57, 62, 120–128, 136, 180,
Bioactive compounds, 57, 92, 243, 290, 191–192, 296
292–294, 297 By-products, 6, 11, 21, 23, 73–75, 81, 83, 87,
Bioconversion, 36, 52, 54, 55, 62, 104, 105, 88, 91, 101, 103, 122, 133, 139,
140, 147, 193, 201, 225, 249, 269, 147, 172, 178, 180, 190, 196, 198,
273, 274, 276, 281, 282, 311 207, 208, 212, 218, 225, 243, 244,
Biodegradability, 23, 24, 117, 127, 154, 158, 247–249, 263, 280, 291–295, 318,
160, 163, 164, 216, 218 319, 323
Biodiversity, 13, 22, 25, 26, 28, 29, 33, 38–41,
43, 45, 239–241, 244, 250, 318,
321, 325 C
Bioeconomy, 54, 56, 57, 93, 154, Circular economy (CE), 37, 53–57, 69–71,
239–250, 318 100, 102, 103, 129, 171–184, 191,
Bioenergy, 5, 12, 37, 51–62, 99–105, 115, 199, 200, 208, 232, 242–244,
118, 120, 172–184, 190, 191, 217, 247–250, 288, 319, 323, 325,
239, 242–244, 247, 249, 250, 308, 328, 330
311, 323 Climate change, 1–3, 5, 7–10, 13, 28, 38–40,
Biofuels, 1–14, 36–38, 51, 53, 54, 57, 58, 42, 55, 71, 99–102, 104, 115, 129,
69–74, 76, 84, 86–88, 91, 93, 134, 182, 201, 239, 240, 244
99–102, 104, 105, 109, 110, Conversion technologies, 51–53, 94, 100, 116,
115–129, 145–148, 154, 173, 175, 119–121, 128, 129, 176, 184, 243
176, 179, 182–184, 190–194,
197–199, 201, 214, 217–219, 224,
225, 227, 231, 241–243, 245, 247, E
249, 270, 276–277, 281, 295, 297, Economy, 45, 54, 71, 111, 183, 208, 249, 255,
303–312, 323, 326 289, 325

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 335
Springer Nature Switzerland AG 2024
A. L. Srivastav et al. (eds.), Valorization of Biomass Wastes for Environmental
Sustainability, https://doi.org/10.1007/978-3-031-52485-1
336 Index

Enhancement strategies, 116, 126–129 Microorganisms, 5, 6, 26, 28, 29, 40, 59, 69,
Environmental pollution, 69, 92, 117, 244, 85, 94, 104, 106, 117, 125, 146,
247, 249, 325 160, 179, 190, 196–198, 212–215,
Enzymatic hydrolysis, 51, 73, 82, 84–89, 92, 224, 228, 229, 231, 239, 241, 245,
93, 100, 102, 104–110, 134, 246, 248, 257, 270, 272, 273, 279,
140–149, 182, 197–199, 211, 228, 280, 295–297, 303, 305–307, 309,
277, 295, 303 310, 327
Enzymes, 57, 69, 70, 72–88, 90–94, 100, 102, Myco-degradation, 269–282
104–109, 125, 126, 137, 139–142,
144–148, 162, 180, 181, 190–199,
202, 208, 210, 212–215, 218, N
223–232, 245, 261, 269–282, 293, Nanotechnology, 101, 106–108, 110, 111
294, 296, 297, 303–311

O
F Organic, 22, 57, 69, 105, 115, 142, 173, 190,
Feedstock, 2–8, 11–14, 37, 52, 54, 57–59, 62, 208, 223, 240, 251, 274, 289,
80, 87, 91, 93, 94, 99–104, 305, 318
116–120, 122–126, 129, 137, Organic residues, 290
139–141, 143, 146–149, 160, 161,
166, 171, 172, 177, 178, 180–184,
190, 197–199, 201, 213–216, P
225, 227, 241–243, 247, 249, Pesticides, 21–23, 26, 27, 29, 41, 192, 193,
263, 269, 270, 277, 295, 207, 223, 224, 226, 230, 240, 241,
310, 323 244, 258, 259, 319, 321
Fossil fuel, 1–3, 5–8, 10–14, 22, 36–38, 53, Pharmaceuticals, 89, 92, 164, 167, 196, 244,
55, 60, 71, 72, 88, 99–101, 108, 279, 294
110, 115, 118, 126, 129, 133, Pretreatment, 51–53, 61, 73, 80–86, 90, 93,
159–161, 171–173, 183, 191, 201, 100, 102, 104–110, 115–117, 120,
229, 231, 239–242, 247, 249, 276, 121, 124–129, 133–149, 177, 182,
279, 303, 318, 323 194, 207–219, 228, 276, 277,
279–281, 295, 296, 303
Pretreatment methods, 73, 80, 82–86, 93, 102,
G 105–108, 110, 115–117, 126, 127,
Greenhouse gas (GHG), 1–3, 7, 10–12, 22, 27, 133–149, 177, 207–219, 303
36, 38, 39, 41, 43, 70, 71, 93,
99–102, 110, 118, 133, 161,
172–174, 177, 181–183, 191, R
199–201, 240, 246, 249, 257, 276, Renewable, 1, 33, 51, 87, 99, 117, 133,
318, 329 154, 171, 199, 208, 232, 239,
270, 303
Renewable resources, 71, 73, 91, 101, 102,
L 104, 191, 239, 240, 242, 246, 247
Lignocelluloses, 103, 107–108, 118, 139, 143, Resource, 21, 51, 69, 100, 125, 154, 172, 189,
145, 195, 197, 208, 210, 269–271, 208, 231, 239, 252, 276, 291,
273, 275, 276, 280–282, 303, 318
304–311, 318
Lignolytic enzymes, 272, 278, 281
S
Solid waste, 4, 55, 57, 58, 62, 75, 115,
M 119–120, 143, 174, 190, 224, 225,
Management, 22, 52, 69, 101, 134, 179, 189, 232, 252, 258, 269, 275, 287, 290,
208, 231, 240, 253, 288, 318 291, 304, 317
Index 337

Sustainability, 13, 27, 40, 42, 43, 45, 57, 58, V


71, 74, 87, 94, 157, 161, 175, 177, Valorization, 62, 69–94, 102, 107,
183, 198, 200, 208, 240, 243–246, 129, 134, 147,
263, 278, 305, 308, 318, 320, 321, 319, 328
324–326, 330 Value-added products, 55–57, 62, 73,
Sustainable, 22, 51, 69, 99, 129, 145, 154, 76, 88–89, 91–94, 134, 138,
172, 189, 208, 231, 239, 252, 270, 147, 148, 171, 226, 239, 243,
288, 320 249, 263, 269,
Sustainable agriculture, 21, 22, 31, 38, 41, 287–298, 319
43–45, 244, 250, 261, 317–330
Sustainable bioenergy, 58, 61
Sustainable waste management, 55, 94, 192, W
200, 308 Waste, 1, 21, 51, 70, 99, 115, 133, 156, 172,
189, 207, 223, 239, 251, 269, 287,
304, 317
T Waste management, 21, 36, 38, 51–53, 57,
Technologies, 28, 51, 74, 100, 115, 147, 153, 69, 81, 134, 179, 193, 199, 201,
173, 190, 208, 223, 239, 252, 274, 208, 223–232, 239, 245, 247,
295, 311, 324 258, 261, 265, 298, 305, 306,
Thermochemical conversation, 9, 10, 71, 74, 319, 323
94, 120, 122–124, 175, 176, 182, Waste treatment, 51, 58, 223–225, 228,
209, 242 231, 264
Three-dimensional (3D) printing, 153–167 White rot basidiomycetes, 272
Trash conversion, 323

Z
U Zero hunger, 14, 329
Unsaturated fattyacids, 293

You might also like