You are on page 1of 267

Suhaib 

A. Bandh   Editor

Strategizing
Agricultural
Management
for Climate Change
Mitigation and
Adaptation
Strategizing Agricultural Management for Climate
Change Mitigation and Adaptation
Suhaib A. Bandh
Editor

Strategizing Agricultural
Management for Climate
Change Mitigation and
Adaptation
Editor
Suhaib A. Bandh
Department of Higher Education
Government of Jammu and Kashmir
Srinagar, India

ISBN 978-3-031-32788-9    ISBN 978-3-031-32789-6 (eBook)


https://doi.org/10.1007/978-3-031-32789-6

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Switzerland AG 2023
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
Dedicated
To
Prof. Azra N. Kamili
Preface

Agriculture is a sector that is deeply affected by climate change. As the world’s


population continues to grow and the demand for food increases, climate change
will pose significant challenges to the sustainability of agricultural systems.
Improvements in agriculture over the past 50 years have made it feasible to meet the
food, feed, and fiber needs of the world’s largest human population. However, due
to widespread urbanization, severe land degradation, and climate change, it is dif-
ficult to keep the supply for the ever-increasing human population. Given its nega-
tive impact on agricultural output, the latter is arguably the greatest food security
danger of the twenty-first century. Agricultural output has been significantly
impacted by global climate change, and agriculture is also a major contributor to
climate change by raising atmospheric greenhouse gas concentrations. To strike a
healthy balance between agricultural output and associated climate effects, there
has never been a greater need for innovative approaches to agricultural technology
and management. To meet the challenges of maintaining the continuous rise in agri-
cultural production, bolstering the resilience of crop production to climate change,
and guaranteeing success in climate change mitigation, adjustments in agriculture
management are necessary. However, the success of agricultural management-based
adaptation and mitigation of climate change, demands the development of straight-
forward, economical, and highly flexible approaches. As a result, planning for agri-
cultural administration is crucial for dealing with climate change. A significant
increase in the input from the scientific community in this research area would play
a key role in understanding how strategizing agricultural management can help in
adapting to climate change and mitigating its effects on agricultural production.
With careful planning and innovative techniques, we can turn agriculture into a
solution rather than a problem. By adopting sustainable farming methods like agro-
forestry, crop rotation, cover cropping, and reducing tillage intensity, farmers can
sequester carbon in the soil while improving their yields and increasing resilience
against extreme weather events. Furthermore, promoting local food systems that
reduce transportation emissions and supporting regenerative farming practices that
prioritize biodiversity conservation can help mitigate the negative impacts of agri-
culture on the environment. Improving water management through measures like

vii
viii Preface

conservation agriculture can have a significant impact on mitigating climate change


in agriculture. These strategies are achievable and can be implemented across a
wide range of agricultural landscapes to mitigate the impacts of climate change.
While these measures might require investments in technologies and training, they
are feasible and have the potential to deliver benefits to both the environment and
the farmers. Therefore, such approaches will be needed to address the challenges
that climate change presents to the agricultural sector, ensuring its continued pro-
ductivity in the future.

Pulwama, India Suhaib A. Bandh


Acknowledgements

If every project has its secret inspiration or at least its one motivation, here that is the
PEACE of my life.

In the name of Allah, the Most Gracious, the Most Merciful. May the praise of
Allah, in the highest of assemblies, and His peace, safety and security, both in this
world and the next, be on Prophet Mohammad (peace be upon Him), the best of
humankind, the most respectable personality for whom Allah created the whole
universe and the seal of the Prophets and Messengers. I am highly thankful to Allah,
who provided me with the courage and guidance to undertake and complete this
project with his great mercy and benevolence.
First and foremost, I would like to thank all my teachers who held my finger to
tread the learning path and enabled me to compile a book. I deeply thank them for
the advice and encouragement, which guided my personal and professional devel-
opment. Publication of a research article, review article or book requires the efforts
of many people besides the authors. I wish to express my special appreciation to
Springer’s editorial and production staff including Herbert Moses and Henry
Rodgers for their excellent and efficient work. In particular, I would like to thank
Aaron Schiller, Senior Editor Springer Nature, for his unwavering confidence in me.
They were there whenever needed and they supervised the production of this project
with commendable attention to all its minute and vivid details. Inevitably, a book of
this type relies heavily on previously published work. I want to thank all the copy-
right holders for granting permission to publish original diagrams and data. My
special thanks go to Keith McNeill, Vincent Happy Ogwugwa, Mahroz Hussain,
Abdullah Kavini Rad, Dr. Muhammad Irfan, Charles Anukwonke, Vahid Karimi,
Dr. Muhammad Ziaul Hoque, Naser Valizadeh, Pritha Datta, Elame Fouad, and Dr.
Muhammad Adil for their valuable contributions to this book. I wish to extend my
appreciation to all the people who assisted me individually in completing this proj-
ect. I am extremely grateful to all the reviewers who provided their timely inputs to
improve the quality of the chapters. My special thanks are due to all those who have
directly or indirectly worked for the successful completion of this project. Finally,

ix
x Acknowledgements

but most importantly, I wish to extend my appreciation to my family for their


patience and encouragement during the period of its compilation. I owe a debt of
gratitude to all of them for their patient forbearance and unwavering support.

Pulwama, India Suhaib A. Bandh


Contents

 1 Nitrogen
 Fertilizer Application Techniques to Reduce
Nitrous Oxide Emissions ������������������������������������������������������������������������    1
Vincent Happy Ogwugwa and Suhaib A. Bandh
 2 Rice
 Production Technologies in Reducing Methane
Gas Emissions for Sustainable Environment����������������������������������������   11
Hamna Bashir, Irshad Bibi, Nabeel Khan Niazi, Abdul Qadeer,
Shumaila Zaman, Ayesha Farzand, Muhammad Mahroz Hussain,
and Muhammad Ashir Hameed
 3 Manure
 Management to Reduce Methane Emissions��������������������������   29
Abdullah Kaviani Rad, Hassan Etesami, Angelika Astaikina,
and Rostislav Streletskii
 4 Crop
 Residue Incorporation to Enhance Soil Health
in the Rice–Wheat System����������������������������������������������������������������������   47
Hamna Bashir, Waqas Mohy-Ud-Din, Zahoor Mujdded Choudary,
Muhammad Mahroz Hussain, and Muhammad Ashir Hameed
 5 Promoting
 Energy Crops to Replace Fossil Fuel Use ��������������������������   69
Muhammad Irfan, Liu Xianhua, Asia Shauket, Muhammad Jafir,
Adeel Ahmad, Samina Jam Nazeer Ahmad, and Jam Nazeer Ahmad
 6 Changes
 in the Agriculture Sector That Are Essential
to Mitigate and Adapt to Climate Changes ������������������������������������������   89
Enohetta B. Tambe, Charles C. Anukwonke,
Iheoma E. Mbuka-Nwosu, and Chinedu I. Abazu
 7 Adaptation
 and Maladaptation to Climate Change:
Farmers’ Perceptions������������������������������������������������������������������������������  113
Vahid Karimi, Masoud Bijani, Zeynab Hallaj, Naser Valizadeh,
Negin Fallah Haghighi, and Mandana Karimi

xi
xii Contents

 8 Farmers’
 Perception of Climate Change in Climatically
Vulnerable Ecosystem of Bangladesh����������������������������������������������������  133
Foyez Ahmed Prodhan, Muhammad Ziaul Hoque,
Md. Safiul Islam Afrad, Md. Enamul Haque, Minhaz Ahmed,
Md. Humayun Kabir, Md. Sadekur Rahman, and Naima Sultana
 9 Pest
 and Disease Management Under Changing Climate��������������������  149
Yaser Biniaz, Naser Valizadeh, Farshad Hemmati,
and Alireza Afsharifar
10 C
 limate Change Adaptation Through Agroforestry:
Empirical Evidence from Indian Eastern Himalayan Foothills����������  167
Pritha Datta and Bhagirath Behera
11 Policy
 Framework to Introduce Climate-­Smart Agriculture��������������  183
Fatemeh Fathi, Naser Valizadeh, Samira Esfandyari Bayat,
and Khadijeh Bazrafkan
12 Technological
 and Managerial Innovation in Agriculture
to Ensure Food Security Under Climate Change����������������������������������  207
Fouad Elame, Hayat Lionboui, and Mohammed Behnassi
13 Oyster
 Farming and a Worldwide Referendum on Global
Carbon Fee-and-Dividend����������������������������������������������������������������������  221
Keith McNeill
14 Climate
 Change Impact Modeling on Citrus Yield������������������������������  233
Fouad Elame, Youssef Chebli, Hallam Jamal, and Lionboui Hayat
15 Impact
 of Climate Change on Environmental Fate
and Ecological Effects of Pesticides��������������������������������������������������������  247
Muhammad Adil, Ghazanfar Abbas, Rabia Naeem Khan,
and Faheem Abbas

Index������������������������������������������������������������������������������������������������������������������  265
Chapter 1
Nitrogen Fertilizer Application Techniques
to Reduce Nitrous Oxide Emissions

Vincent Happy Ogwugwa and Suhaib A. Bandh

Abstract  Mankind has developed a technique to generate synthetic reactive nitro-


gen, which serves as a fertilizer to enhance food production. However, once reactive
nitrogen molecules are formed, they become highly mobile and can persist in the
environment for extended periods, leading to various negative consequences. One
such consequence is the increased emission of nitrous oxide (N2O), a long-lived
radiatively active greenhouse gas (GHG) with a molecular heat trapping effect
approximately 310 times stronger than other gases, posing a significant environ-
mental challenge. To address this issue, optimizing nitrogen (N) fertilization
becomes crucial. The goal is to match the supply of nitrogen from fertilizers with
the crop’s demand, thereby reducing excess soil nitrogen. By achieving this bal-
ance, the production of soil N2O can also be minimized. Several strategies can be
employed to achieve this, such as using slow-release fertilizers that gradually release
nutrients into the soil. Additionally, the use of chemical urease inhibitors (UI) and
nitrification inhibitors (NI) can slow down the conversion of urea to NH4+, further
reducing nitrogen loss. Another effective approach is to adopt a split application
method, where fertilizer is applied multiple times throughout the crop cycle. This
strategy aims to synchronize fertilizer application with the rapid nitrogen demand of
the plants, thereby minimizing N2O emissions. Furthermore, foliar nitrogen fertil-
izer application can be employed, allowing the active absorption of nitrogen into the
interior of the leaf blade through plasmodesmata, hydrophilic pores in the waxy
cuticle of the leaf surface, and stomata distributed on the leaf surface. These mecha-
nisms enable the efficient absorption of available nutrients. Moreover, the applica-
tion of nanotechnology offers a promising solution by reducing the reactivity of
nutrient inputs into the agricultural system without compromising productivity,
holding a great potential for sustainable agriculture.

Keywords  Fertilizer application · N2O emission · Agriculture and nutrient

V. H. Ogwugwa (*)
Department of Microbiology, University of Lagos, Lagos, Nigeria
S. A. Bandh
Department of Higher Education, Government of Jammu and Kashmir, Srinagar, India

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


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_1
2 V. H. Ogwugwa and S. A. Bandh

1 Introduction

In 2015, the UN outlined 17 Sustainable Development Goals (SDGs) (Wang et al.,


2022). According to studies, there may be links between various aims that are
unbreakable, consistent, etc. Goal 2 focuses on agricultural productivity, which
aims to end world hunger. Nitrogen fertilizer is a crucial agricultural input for grain
growth, but it also has many negative environmental effects (Shakoor et al., 2021).
Excess nitrogen fertilizer use may cause significant nitrogen losses, which contrib-
ute to eutrophication of freshwater and atmospheric pollution, which is very unfa-
vorable for mitigating climate change, maintaining ecosystem health, and improving
water quality. Too little nitrogen fertilizer use may result in low yields, which makes
it very unfavorable for ending poverty and hunger, and improving human health.
The secret to encouraging the implementation of SDGs related to nitrogen fertilizer
is to optimize the application of nitrogen fertilizer (Bai & Gao, 2021).
It is well established that nitrogen (N) has a significant influence in biodiversity
loss, air pollution (through NOx and NH3 emissions and indirectly on O3), and cli-
mate change (via N2O emissions and aerosols and indirectly on CO2, CH4, and O3
emissions). As a result of their strong connections to food security, N use and cli-
mate change are two highly essential concerns to take into account (Maaz et al.,
2021). Since 1980, China’s grain production has increased by about 70%, and it has
also grown to be one of the world’s top emitters of greenhouse gases (GHGs) due to
a threefold increase in the usage of N fertilizer in agriculture (Liu et al., 2021). Due
to the fast-rising fertilizer N use since 1980, the effects of N fertilizer use on crop
output, greenhouse gas (GHG) emissions, and climate change will become more
and more significant (Bandh et al., 2021; Bandh, 2022a). The cycling of nitrogen in
terrestrial ecosystems may be accelerated by climate change, particularly climatic
warming. As a result of fertilizer application, additional N2O and other GHG gases
may be released from both croplands and nonarable soils, thus accelerating the rate
of climate change (Lan et al., 2021). Additionally, carbon (C) sequestration in ter-
restrial ecosystems may be impacted by N cycling; this must be considered when
forecasting future CO2 uptake. Therefore, it is important to take climate change
seriously, especially in emerging economies, and to consider the relationship
between N fertilizer application, GHG emissions, and climate change (Shakoor
et al., 2021).

2 History of Fertilizer Application

Since hundreds of years ago, organic and mineral fertilizers have been applied to
grasslands (Ashekuzzaman et  al., 2021). Paddock manuring is the most notable
instance of the use of organic fertilizer, and it has been done in many areas for mil-
lennia. The movement of nutrients by livestock on pasture forms the basis of pad-
dock manuring (Garcia et  al., 2021). Small transportable enclosures are used to
1  Nitrogen Fertilizer Application Techniques to Reduce Nitrous Oxide Emissions 3

house livestock that grazes on pasture during the day. The majority of the nutrients
ingested by animals in other pastures are released as feces and urine at night inside
this enclosure. The enclosure is advanced to the sections that need fertilizing after a
set amount of time, often 1 week. To take advantage of the fertilizer’s impact on
biomass output and fodder quality in later years, such fertilized regions were tradi-
tionally most frequently employed for haymaking (Ashekuzzaman et  al., 2021
Garcia et al., 2021).
Humans have been known to transport and apply nutrients in numerous areas of
Europe for ages. For instance, throughout the seventeenth and twentieth centuries,
hay was transported from subalpine grasslands down the valleys across a distance of
10–15 km in the Giant Mountains (Krkonoe, Riesengebirge) situated in the border-
land between the Czech Republic and Poland. This hay was fed to the cows and
goats maintained in the valleys’ cow huts. In close proximity to the farm homes, the
grassland or arable land received an application of the organic fertilizer nutrients
derived from that hay. As a result, grasslands at lower altitudes saw an increase in
biomass output, while grasslands in the subalpine vegetation zone experienced
nutrient depletion.

3 Effects of N Fertilizer Use on the Environment

The effects of nitrogen fertilizer use on water and air pollution have drawn more
attention recently (Martínez-Dalmau et al., 2021). These environmental worries are
valid and will become more significant in the coming years. Numerous nations have
enacted restrictions on fertilizer use, and further initiatives to do so are likely. In
locations where manure and N fertilizer are currently used at rather high levels, such
interventions might be necessary. However, the low rates of use combined with
increasing population pressure on the limited land resources in the majority of the
developing world force expanded fertilizer use to fulfill rising food needs (Kim
et al., 2021).

3.1 Emissions Generated by Fertilizer Application

3.1.1 Nitrous Oxide

Nitrous oxide (N2O), despite being largely inactive in the troposphere, absorbs
infrared light (de Vries, 2021). In the stratosphere, it can also photochemically
break down to create nitric oxide, which can interact with ozone. Since roughly
1960, the concentration of N2O has been observed at an increasing rate. Currently,
the atmosphere’s N2O concentrations are rising at a rate of roughly 0.3% per year
(Cao et al., 2021). There are around 1500 Tg (1012 g/Terra gram) of N2O in the
atmosphere. Nearly 90% of the emissions come from soils via biologically driven
4 V. H. Ogwugwa and S. A. Bandh

nitrification and denitrification processes. In each of these processes, N2O emission


might be seen as a leakage of intermediate products (Wei et al., 2022). While soils
play a significant role in the production of N2O, they also act as a sink. The forma-
tion of N2O from nitrification is a topic of significant research because worries about
N2O as an atmospheric pollutant have increased recently (Ge et al., 2021). The N2O
emissions are extremely variable as a result of this process, which is to be expected
given the wide range of variables influencing the reaction, such as oxygen avail-
ability, water content, temperature, soil pH, organic matter, and the presence of
plants. Emissions are typically proportional to the amount of nitrifiable nitrogen in
the soil.

3.1.2 Nitric Oxide and Nitrous Oxide

Though they are regarded to not be as significant in absorbing light energy as N2O,
nitric oxide and nitrous oxide (together referred to as NOx) are involved in signifi-
cant atmospheric processes (Han et al., 2021). Ozone (O3) and hydroxyl radicals are
created when NO oxidizes CH4 and CO at levels that are high enough (>10 ppt). The
primary purifying component in the troposphere is hydroxyl radicals. Ozone (O3)
and hydroxyl radicals may both experience net losses due to lower NO concentra-
tions. According to Han et  al. (2021), NOx can destroy ozone, especially in the
stratosphere, and NO2 forms nitric acid when it combines with hydroxyl ions in a
photochemically stimulated reaction. Estimates of emissions are primarily based on
laboratory studies, and the production of NOx from soils has not been explored to
the same extent as N2O. The processes of nitrification, denitrification, chemo deni-
trification, and perhaps photolysis of nitrite are used to create the gases (Harris
et  al., 2021). The primary contributors seem to be nitrification mechanisms.
Although it is impossible to make judgments about how much N fertilizers contrib-
ute to NOx emissions, it is most certainly true; the absolute amounts, however, are
quite unpredictable.

4 Climate Change and Its Impact due to Release


of Greenhouse Gases (GHGs)

According to the United Nations Framework Convention on Climate Change


(UNFCCC), climate change has affected food security and human health in many
regions and has been a global issue. One of the main causes of climate change,
according to experts, is rising worldwide GHG emissions (Oreggioni et al., 2021).
As a result, the UN gave the subject of climate change high priority and included it
as one of the SDGs. It must work to create a resource-conserving, environmentally
friendly society and strengthen the nation’s ability to reduce GHG emissions.
1  Nitrogen Fertilizer Application Techniques to Reduce Nitrous Oxide Emissions 5

5 Nitrogen Fertilizer Application Techniques to Reduce


Nitrous Oxide Emissions

In order to maximize both economic and ecological needs, increasing N fertilizer


use efficiency through novel methods and technology is a challenge. By using spe-
cialized application methods, such as applying NH4-dominated fertilizer, as well as
novel fertilizer technologies, such as coating or the use of nitrification and urease
inhibitors, it is possible for farmers to utilize less split N applications (Aliyu et al.,
2021). These methods therefore attempt to reduce unproductive N losses and, con-
sequently, negative environmental effects brought on by NO3 leaching and/or gas-
eous N losses while increasing productivity through an optimized fertilizer
usage rate.

5.1 Precision Agriculture (Site-Specific Application)

Precision agriculture (PA) technology has advanced quickly over the past few
decades, particularly with regard to arable crops, with the aim of facilitating the
handling of production tools and resources in an economically viable and sustain-
able manner through site-specific management. PA could be defined more broadly
as the application of information technology across the board in agriculture (Nyaga
et al., 2021). The accurate calculation of the necessary N substitution after harvest
on a per-plot basis is one method for handling recycled N in farms in an environ-
mentally beneficial manner. It is necessary to periodically check the soil’s nutrient
level due to the unequal distribution of nutrients throughout farms. Such a method,
nevertheless, does not currently take field heterogeneity into account. Although the
challenge of applying fertilizer precisely to farms is not new, the spatial component
is now more widely acknowledged as site-specific application technology advances
quickly. Site-specific slurry application is seen as a significant contribution to envi-
ronmentally friendly and efficient nutrient use on grassland when done in accor-
dance with the best practice for nutrient management (Bandh, 2022b).

5.2 Use of Nitrification Inhibitors (NI)

The biological nitrification and denitrification processes are the primary sources of
N2O in soils. As a non-obligate intermediate during autotrophic nitrification, the
hydroxylamine oxidoreductase (HAO) in aerobic soil compartments produces the
gas N2O during the oxidation of hydroxylamine. As the first stage of nitrification,
NH4’s transformation into NO2 is slowed down by nitrification inhibitors (NI)
(Wang et al., 2021). By preserving the immobile NH4 in the soil for a set amount of
time, this delay lowers the quantity of NO3 that can be readily leached under specific
6 V. H. Ogwugwa and S. A. Bandh

conditions. Additionally, NIs lessen the internal N mineralization of the soil, reduc-
ing the possibility of soil NO3 surpluses that could be lost.

5.3 Use of Urease Inhibitor (UI)

One of the most popular nitrogen fertilizers is urea, which makes up more than half
of the annual consumption of chemical nitrogen fertilizers. Urea is hydrolyzed
when applied to soil, and the NH3 that results from this process can be volatilized,
resulting in significant financial loss and environmental pollution. Urease inhibitor
prolongs the duration that urea diffuses at fertilizer application sites by delaying the
water dissolution of urea (Klimczyk et al., 2021). This will decrease the density of
NH4+ and NH3 in the soil and lessen the amount of ammonia lost through volatiliza-
tion. After 30 years of development, there are more than 100 different types of ure-
ase inhibitors. The predominant varieties include humic acid, quinines, acidamide,
polyacid, polyphenol, etc.

5.4 Slow-Release/Controlled-Release Fertilizer

The mismatch between the needs of crops and the frequency and intensity with
which fertilizer releases its nutrients is one of the causes of low fertilizer use effi-
ciency. By adjusting the water solubility of regular fertilizer, slow-release/
controlled-­release fertilizer is created. By enhancing the fertilizer, itself, which
aligns the release time and intensity with the needs of crops, the nitrogen release is
effectively controlled or postponed (Dong et al., 2021). This technique can balance
the supply of nutrients with the demand from crops, thereby increasing the yield. It
is thought to be the simplest and fastest approach to reduce fertilizer loss and its
release of gaseous chemicals into the environment. Slow-release fertilizer mainly
delays the release of the nutrients and extends the fertilizer effect period. Slow-­
release fertilizer combines acceleration and delay of the nutrient release from the
fertilizers; it can control the speed of nutrient supply.

5.5 Biofertilizer

Biofertilizers are useful, living microbes. These include bacterial phosphate solubi-
lizers like Pseudomonas and Bacillus, as well as blue green algae, Rhizobium,
Azotobacter, Azotospirillum, and fungal mycorrhizae. Organic matter is broken
down by microorganisms into simple chemicals that give plants vital nutrients, raise
soil fertility, preserve the soil’s natural habitat, and boost crop output. The success-
ful application of biofertilizers depends on the preparation, storage, and application
1  Nitrogen Fertilizer Application Techniques to Reduce Nitrous Oxide Emissions 7

process. Short shelf life, temperature sensitivity, and storage desiccation issues are
drawbacks in use (Sun et al., 2021). The potential for covering biofertilizer prepara-
tions with polymeric nanoparticles to produce formulations resistant to desiccation
has been explored. One method for storing and distributing microorganisms through
liquid formulations is the water-in-oil emulsion. The oil prevents the water from
evaporating by trapping it around the bacteria. For microorganisms that are suscep-
tible to desiccation, this is very beneficial. By adding chemicals to the oil and/or
aqueous phases, water-in-oil emulsions also enhance cell viability and release kinet-
ics. However, one of the main concerns is sedimentation during storage. By thicken-
ing the oil phase during storage, hydrophobic silica nanoparticles decreased cell
sedimentation and increased cell survival.

5.6 Application of Nanotechnology in Delivery of Fertilizer

Massive volumes of fertilizer in the form of urea, nitrate, or phosphate compounds,


as well as ammonium salts, have significantly enhanced food production, but they
also have many negative impacts on the good soil microflora. Due to runoff, the
majority of fertilizers are not accessible to plants and may result in environmental
damage if used in excess. This issue can be resolved with nanomaterial-coated fer-
tilizers (Rakhimol et al., 2021). As nanoparticles retain the fertilizer more tightly
from the plant due to their higher surface tension than ordinary surfaces, nanomate-
rials may help slow down the release of fertilizers. Additionally, nanocoatings shield
bigger particles’ surfaces. In order to limit nutrient losses and undesired nutrient
interactions with microbes, water, and air, nanofertilizers balance the release of fer-
tilizer nitrogen and phosphorus with the uptake of the plant (Singh et al., 2021).
Utilizing nanofertilizer will increase the plants’ ability to absorb nutrients from the
soil. After nutrient absorption by fungi or bacteria, nanosilica-encapsulated nanofer-
tilizer can form a binary layer on the cell wall, preventing infections and enhancing
plant growth under high temperature and humidity as well as disease resistance.
Due to silicon dioxide nanoparticles’ ability to enhance seedling growth and root
development, silicon-based fertilizers are used to promote plant resilience.

6 Conclusions and Future Perspectives

We looked at current approaches for applying nitrogen fertilizer to cut down on


nitrous oxide emissions. Nitrous oxide emissions can be reduced by using nitrifica-
tion inhibitors, urease inhibitors, site-specific applications, slow-release/controlled-­
release fertilizer, biofertilizer, and nanotechnology in fertilizer delivery. However, it
is important to consider the potential trade-off with higher direct emissions.
Although anaerobic digestion has proven successful in lowering NH3 and N2O
emissions, more study is required to determine the effects of off-farm inputs.
8 V. H. Ogwugwa and S. A. Bandh

Injection is preferred on well-drained soil, whereas surface spreading followed by


quick assimilation is preferred on damp soil. Appropriate manure application meth-
ods can reduce direct and indirect N2O emissions. The timing of manure application
prevents excessive N2O fluxes and decreases NO3 leaching. In general, N2O emis-
sions can be reduced by applying only as much manure as is necessary to meet the
N requirements of the crop. Application rate is also crucial. In tests, nitrification
inhibitors can lower direct and indirect N2O emissions; however, it is not apparent
if this mitigation is economically viable at the farm size. Therefore, additional
investigation is required.

References

Aliyu, G., Luo, J., Di, H.  J., Liu, D., Yuan, J., Chen, Z., He, T., & Ding, W. (2021). Yield-­
scaled nitrous oxide emissions from nitrogen-fertilized croplands in China: A meta-analysis
of contrasting mitigation scenarios. Pedosphere, 31(2), 231–242. https://doi.org/10.1016/
S1002-­0160(20)60074-­1
Ashekuzzaman, S. M., Forrestal, P., Richards, K. G., Daly, K., & Fenton, O. (2021). Grassland
phosphorus and nitrogen fertiliser replacement value of dairy processing dewatered
sludge. Sustainable Production and Consumption, 25, 363–373. https://doi.org/10.1016/j.
spc.2020.11.017
Bai, Y., & Gao, J. (2021). Optimization of the nitrogen fertilizer schedule of maize under drip
irrigation in Jilin, China, based on DSSAT and GA. Agricultural Water Management, 244,
106555. https://doi.org/10.1016/j.agwat.2020.106555
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Cao, Y., Wang, R., Peng, J., Liu, K., Chen, W., Wang, G., & Gao, X. (2021). Humidity enhanced
N2O photoacoustic sensor with a 4.53 μm quantum cascade laser and Kalman filter.
Photoacoustics, 24, 100303. https://doi.org/10.1016/j.pacs.2021.100303
de Vries, W. (2021). Impacts of nitrogen emissions on ecosystems and human health: A mini
review. Current Opinion in Environmental Science and Health, 21. https://doi.org/10.1016/j.
coesh.2021.100249
Dong, D., Li, J., Ying, S., Wu, J., Han, X., Teng, Y., Zhou, M., Ren, Y., & Jiang, P. (2021). Mitigation
of methane emission in a rice paddy field amended with biochar-based slow-release fertilizer.
Science of the Total Environment, 792, 148460. https://doi.org/10.1016/j.scitotenv.2021.148460
Garcia, J. S., Anderson, K. E., Guard, J. Y., Gast, R. K., & Jones, D. R. (2021). Impact of organic
dairy cattle manure on environmental and egg microbiology of organic free-range laying hens.
Journal of Applied Poultry Research, 30(4). https://doi.org/10.1016/j.japr.2021.100189
Ge, P., Chen, M., Cui, Y., & Nie, D. (2021). The research progress of the influence of agricultural
activities on atmospheric environment in recent ten years: A review. Atmosphere, 12(5). https://
doi.org/10.3390/atmos12050635
Han, Z., Wang, J., Xu, P., Li, Z., Liu, S., & Zou, J. (2021). Differential responses of soil nitrogen-­
oxide emissions to organic substitution for synthetic fertilizer and biochar amendment in a sub-
tropical tea plantation. GCB Bioenergy, 13(8), 1260–1274. https://doi.org/10.1111/gcbb.12842
1  Nitrogen Fertilizer Application Techniques to Reduce Nitrous Oxide Emissions 9

Harris, E., Diaz-Pines, E., Stoll, E., Schloter, M., Schulz, S., Duffner, C., Li, K., Moore, K. L.,
Ingrisch, J., Reinthaler, D., Zechmeister-Boltenstern, S., Glatzel, S., Brüggemann, N., & Bahn,
M. (2021). Denitrifying pathways dominate nitrous oxide emissions from managed grassland
during drought and rewetting. Science Advances, 7(6). https://doi.org/10.1126/sciadv.abb7118
Kim, D.-G., Grieco, E., Bombelli, A., Hickman, J. E., & Sanz-Cobena, A. (2021). Challenges and
opportunities for enhancing food security and greenhouse gas mitigation in smallholder farm-
ing in sub-Saharan Africa. A review. Food Security, 13(2), 457–476. https://doi.org/10.1007/
s12571-­021-­01149-­9
Klimczyk, M., Siczek, A., & Schimmelpfennig, L. (2021). Improving the efficiency of urea-based
fertilization leading to reduction in ammonia emission. Science of the Total Environment, 771,
145483. https://doi.org/10.1016/j.scitotenv.2021.145483
Lan, T., Zhang, H., Han, Y., Deng, O., Tang, X., Luo, L., Zeng, J., Chen, G., Wang, C., & Gao,
X. (2021). Regulating CH4, N2O, and NO emissions from an alkaline paddy field under
rice–wheat rotation with controlled release N fertilizer. Environmental Science and Pollution
Research International, 28(14), 18246–18259. https://doi.org/10.1007/s11356-­020-­11846-­1
Liu, Y., Ge, T., van Groenigen, K.  J., Yang, Y., Wang, P., Cheng, K., Zhu, Z., Wang, J., Li, Y.,
Guggenberger, G., Sardans, J., Penuelas, J., Wu, J., & Kuzyakov, Y. (2021). Rice paddy soils
are a quantitatively important carbon store according to a global synthesis. Communications
Earth and Environment, 2(1). https://doi.org/10.1038/s43247-­021-­00229-­0
Maaz, T. M., Sapkota, T. B., Eagle, A. J., Kantar, M. B., Bruulsema, T. W., & Majumdar, K. (2021).
Meta-analysis of yield and nitrous oxide outcomes for nitrogen management in agriculture.
Global Change Biology, 27(11), 2343–2360. https://doi.org/10.1111/gcb.15588
Martínez-Dalmau, J., Berbel, J., & Ordóñez-Fernández, R. (2021). Nitrogen fertilization. A review
of the risks associated with the inefficiency of its use and policy responses. Sustainability,
13(10). https://doi.org/10.3390/su13105625
Nyaga, J.  M., Onyango, C.  M., Wetterlind, J., & Söderström, M. (2021). Precision agriculture
research in sub-Saharan Africa countries: A systematic map. Precision Agriculture, 22(4),
1217–1236. https://doi.org/10.1007/s11119-­020-­09780-­w
Oreggioni, G.  D., Ferraio, M., Crippa, M., Muntean, M., Schaaf, E., Guizzardi, D., Solazzo,
E., Duerr, M., Perry, M., & Vignati, E. (2021). Climate change in a changing world: Socio-­
economic and technological transitions, regulatory frameworks and trends on global green-
house gas emissions from EDGAR v.5.0. Global Environmental Change, 70, 102350. https://
doi.org/10.1016/j.gloenvcha.2021.102350
Rakhimol, K.  R., Thomas, S., & Nandakumar Kalarikkal, J.  K. (2021). Nanotechnology in
controlled-­release fertilizers. In F.  B. Lewu, T.  Volova, & R.  K. R.  Sabu Thomas (Eds.),
Controlled release fertilizers for sustainable agriculture (pp. 169–181). Academic. https://doi.
org/10.1016/B978-­0-­12-­819555-­0.00010-­8
Shakoor, A., Shahzad, S. M., Chatterjee, N., Arif, M. S., Farooq, T. H., Altaf, M. M., Tufail, M. A.,
Dar, A. A., & Mehmood, T. (2021). Nitrous oxide emission from agricultural soils: Application
of animal manure or biochar? A global meta-analysis. Journal of Environmental Management,
285, 112170. https://doi.org/10.1016/j.jenvman.2021.112170
Singh, S. K., Patra, A., Verma, Y., Chattopadhyay, A., Rakshit, A., & Kumar, S. (2021). Potential
and risk of nanotechnology application in agriculture vis-à-vis nanomicronutrient fertilizers.
In A. Rakshit, S. Singh, P. Abhilash, & A. Biswas (Eds.), Soil science: Fundamentals to recent
advances. Springer. https://doi.org/10.1007/978-­981-­16-­0917-­6_26
Sun, H., Zhang, Y., Yang, Y., Chen, Y., Jeyakumar, P., Shao, Q., Zhou, Y., Ma, M., Zhu, R.,
Qian, Q., Fan, Y., Xiang, S., Zhai, N., Li, Y., Zhao, Q., & Wang, H. (2021). Effect of bio-
fertilizer and wheat straw biochar application on nitrous oxide emission and ammonia vola-
tilization from paddy soil. Environmental Pollution, 275, 116640. https://doi.org/10.1016/j.
envpol.2021.116640
Wang, X., Bai, J., Xie, T., Wang, W., Zhang, G., Yin, S., & Wang, D. (2021). Effects of bio-
logical nitrification inhibitors on nitrogen use efficiency and greenhouse gas emissions in
10 V. H. Ogwugwa and S. A. Bandh

a­ gricultural soils: A review. Ecotoxicology and Environmental Safety, 220, 112338. https://doi.
org/10.1016/j.ecoenv.2021.112338
Wang, Y., Lu, Y., Yuan, J., & He, G. (2022). Evaluating the risks of nitrogen fertilizer-related grain
production processes to ecosystem health in China. Resources, Conservation and Recycling,
177. https://doi.org/10.1016/j.resconrec.2021.105982
Wei, Z., Shan, J., Well, R., Yan, X., & Senbayram, M. (2022). Land use conversion and soil mois-
ture affect the magnitude and pattern of soil-borne N2, NO, and N2O emissions. Geoderma,
407, 115568. https://doi.org/10.1016/j.geoderma.2021.115568
Chapter 2
Rice Production Technologies in Reducing
Methane Gas Emissions for Sustainable
Environment

Hamna Bashir, Irshad Bibi, Nabeel Khan Niazi, Abdul Qadeer,


Shumaila Zaman, Ayesha Farzand, Muhammad Mahroz Hussain,
and Muhammad Ashir Hameed

Abstract  Agriculture is the primary contributor for greenhouse gas (GHG), with
rice being a driver to global warming due to large number of GHGs emissions, hav-
ing importance as a staple food for more than half of the world’s population. Global
rice demand to fulfil hunger requirement will enhance GHG emissions due to its
cultivation under paddy rice conditions that have detrimental effects on the environ-
ment results in more GHGs. Agriculture is the major contributor for CH4 emission
under paddy soils, which have increased by methanogen activity resulting in CH4
production, a major end product in the anaerobic food chain (CH4-producing bacte-
ria). Because anaerobic conditions enhance the performance of methanogens, which
leads to the utilization of organic carbon and its conversion into CH4 through a
process called methanogenesis, anaerobic conditions are the biochemical routes of
CH4 generation. Therefore, it is in dire need of time to study the impacts of the
paddy rice system and possible alternative ways to develop critical understanding to
mitigate CH4. In this chapter, production technologies and importance of rice as a
staple food were elucidated. Furthermore, the impacts of different production tech-
nologies to mitigate CH4 production was also discussed critically. However, future
research directions and major research gaps have been identified.

Keywords  Rice · Alternate wetting and drying · Methanogens · Global warming ·


Production technologies

H. Bashir · I. Bibi · N. K. Niazi · A. Qadeer · S. Zaman · A. Farzand


Institute of Soil and Environmental Sciences, University of Agriculture Faisalabad,
Faisalabad, Pakistan
M. M. Hussain (*) · M. A. Hameed
Institute of Soil and Environmental Sciences, University of Agriculture Faisalabad,
Faisalabad, Pakistan
HAM Organics (PVT) Limited, Punjab, Pakistan

© The Author(s), under exclusive license to Springer Nature 11


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_2
12 H. Bashir et al.

1 Introduction

The atmosphere is chemically changed over time due to the energy production by
fossil fuels as they emit greenhouse gases (GHGs) via manufacturing, transporta-
tion, conversion, and combustion. Of all GHGs about 16% is methane, making it a
significant greenhouse gas (Hussain et al., 2021a, c). Despite being steady for mil-
lennia, the amount of methane in the atmosphere has recently increased ~2-fold. In
the absence of oxygen, decomposition of organic molecules causes methane forma-
tion with microbes that acts as a major climate change driver (Din et al., 2022). It is
emancipated from natural and artificial sources, with natural resources causing 40%
of the global methane emissions (Table 2.1), but 60% of these are caused by human
activities (Linquist et al., 2012).
Rice is probably one of the most important grains, as it is consumed by more
than 50% of the population globally (Hussain et al., 2021b). Globally, paddy fields
account for about 11% of the world’s farmland, or 153 million hectares (Wang et al.,
2012). In the next 20 years, demand for rice is expected to grow by 24%, with more
pressure from developing countries to fulfil hunger requirements, being a staple
food (Van Nguyen & Ferrero, 2006). In addition, paddy fields play an important role
in the production of CH4 and N2O and potential sources or sinks of carbon dioxide.
Paddy fields provide roughly 30% of all agricultural CH4 emissions worldwide and
11% of all agricultural N2O emissions (Linquist et al., 2012). The CH4 emissions
from paddy fields are mainly caused by microbial decomposition and the combus-
tion of plant debris and humus (Smith et al., 2008). The organic decomposition in
watered rice crops leads to methane formation, protonation, and transportation (Le
Mer & Roger, 2001; Malla et al., 2022; Bandh et al., 2021, 2023).
For sustainable and profitable rice systems, technologies and methods to offset
greenhouse gas emissions need to be developed. In paddy fields, GHG production is
primarily based on farming practices, but adjustments in management systems have
mitigation potential. Often a practice may be exposed to more than one gas, some-
times through opposing processes, so the net benefit depends on the cumulative
effect of implementation on all gases (Schils et al., 2005).

Table 2.1  Major methane (CH4) emission contributors


Agriculture
(50.63%) Waste (20.61%) Industry (0.10%) Energy (28.65%)
Manure Landfilling of solid Silicone carbide Coal mining activities
management waste production
Rice cultivation Use of solvent and Mineral products Natural gas and oil systems
others
Enteric Waste combustion Metal production Stationary and mobile
fermentation combustion
Other Wastewater Iron and steel Biomass combustion
production
Chemical production
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 13

2 Greenhouse Gas Emission Mechanisms

2.1 CH4 Emissions and Production

Methane is one of the end products of anaerobic food webs in paddy soils due to the
action of methanogens. Methane is produced by anaerobic conditions which is a
biochemical method, as fermentative conditions invigorate the output of methano-
gens that results in the harvesting and conversion of organic carbon into CH4 through
the methanogenesis process (Flynn & Smith, 2010). Due to a range of biological
activities, the redox potential of the rice soil was greatly reduced after being flooded
for a long time (Fig.  2.1). During flooding, methanogenic substrates are mainly
produced by degradation of soil organic compounds. Labile carbon, which is not
taken up by plants, is normally liberated to the soil, where it is rarely transformed
by methanogens into CH4 (Epule et al., 2011).
Three primary processes allow CH4 to be released after methanogenesis. The
dissolved CH4 gas diffusion, boiling loss (discharge of air bubbles mediated via
agricultural reforms or soil fauna), and plant transport to roots via the process of
diffusion and transformation of aerial tissue and gaseous CH4 in the cortex that
consequently liberates CH4 in the atmosphere thru plant micropores are some of
these pathways. Previous studies have reported that more than 90% of the CH4 in
temperate paddy fields is emitted from rice plants (Guo & Zhou, 2007). Nevertheless,
growing tropical rice can also lead to substantial methane emissions, especially in

Fig. 2.1  Factors affecting


methane emissions from
rice field
14 H. Bashir et al.

early seasons when there are high organic amendment applications (van der
Gon, 2000).

3 Agriculture

Annual agricultural methane emissions are estimated at over 3135.7  Mt CO2 eq.
This makes the agricultural sector a large contributor of methane emissions, caused
from human non-point sources. Or we can say that 50.63% of all methane emissions
that are caused by human activities mainly comes from agricultural practices.
Enteric fermentation is the source of 59.84% of all methane emissions, followed by
rice farming, other agricultural pursuits, and waste management. The change in the
percentage of emissions from farms is greater than the change in the percentage of
emissions from other happenings. Since 2000, these emissions have drastically
grown. Moreover, 15.47% of the emissions in this category are produced in China
(Bandh, 2022a, b).

3.1 Source

Agricultural methane emissions are caused through manure management, enteric


fermentation, rice cultivation, and other agricultural practices. The fermentation of
food by microorganisms in the digestive system of animals is termed enteric fer-
mentation (Li et al., 2021). A by-product of this process is methane, which is pro-
duced by the animals’ respiration. Farm animals such as cattle, goats, buffalo,
camels, and sheep are the main source of methane emissions in this sector. Methane
is also produced by enteric fermentation in other domesticated nonruminants, such
as pigs and horses; however, emissions vary greatly from animal to animal. Total
methane emissions from these sources are inversely correlated with the number of
animals, with feed type, quantity, and quality having the most effects. When manure
is stored or processed in lagoons, ponds, or pits, the decomposition process leads to
anaerobic conditions and methane emissions (Bundhoo et al., 2016).
The amount of methane produced by manure depends on the storage method, the
ambient temperature, and the composition of the manure. Higher methane emis-
sions are also feasible in environments with higher room temperature and humidity.
In addition, the feces composition is directly related to the species of the animals
and their diet (Bundhoo et al., 2016). Therefore, the combination of these variables
affects the actual methane emissions caused via manure management. Methane is
produced by the breakdown of oxygen-poor organic molecules in flooded paddy
fields. When paddy fields are flooded, the decomposition of organic matter con-
sumes oxygen in the soil and water gradually. A variety of factors can affect the
amount of methane produced by paddy fields, including the amount of organic mat-
ter present, labile C pool, soil microbiota, and water management (Vaghefi et al.,
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 15

2016). Open combustion of biomass, savannah burning, agricultural waste burning,


and open sweltering after deforestation are examples of sources of agricultural
methane emissions.
Cropping options to reduce GHG emissions while adjusting irrigation patterns
and farming practices, managing organic additives and fertilizers, selecting appro-
priate varieties, and applying cropping systems can all help in reducing greenhouse
gas emissions from paddy fields. In the following sections, the detail of all of these,
as well as options and possibilities in different agroclimatic scenarios, will be
discussed.

4 Management of CH4 via Irrigation Regime Changes

One of the most important factors affecting greenhouse gas emissions is the man-
agement of water resources during rice cultivation. Several water management prac-
tices compared to conventional flooded rice, various midseason drainage intervals,
alternate soil wetting and drying, erratic irrigation, and meticulous irrigation have
been shown to cause reduction in GHG emissions. These techniques can be used in
a variety of soil and climate conditions without affecting crop yield.
Summer Drainage
Midseason drainage is a separate period of irrigation in the agricultural growing
season that is halted. To avoid tillering and reduce unproductive tiller numbers, a
short drainage period of 5–20 days is usually performed before the maximum tiller-
ing phase; the duration is determined by locally accepted traditional practice.
Methane emissions could increase when soil aeration begins due to the release of
CH4 trapped in the soil, followed by a sustained decrease in emissions even if the
field is flooded again. The efficacy of midseason drainage to reduce CH4 varied
significantly (15–59%) due to the additional water available to re-flood the rice soil.
Seasonal drainage improves the soil oxidation state and uptake of nitrogen (Shiratori
et al., 2007). It is achieved due to the decrease in the level of water sprayed, as the
total water level reduction translates into a reduction in CH4 emissions. Because
oxygenation of the soil creates aerobic conditions that are not conducive to metha-
nogen activity, Wassmann et al. (2000b) found that seasonal drainage lowered CH4
discharge by 43%. Timed and controlled midseason drainage seems to be the main
approach to achieve a net reduction in GHG emissions (Wassmann et al., 2000a).
Numerous studies have shown that it works for paddy fields based on total green-
house gas emissions. Methane (CH4)  and N2O emissions advocates midseason
drainage as the most effective method for reducing greenhouse gas emissions, as it
produces a 27% lower global warming potential (GWP)  than typical flooding.
According to Zou et al., the GWPs (CH4 and N2O) of midseason drainage are 42%
and 72% lower, respectively, than typical floods. Because the length and timing of
the drainage season have such a large impact on greenhouse gas emissions, this
management approach can be upgraded to further reduce emissions.
16 H. Bashir et al.

Alternate Wetting and Drying Irrigations


Desiccating and re-flooding a rice field repeatedly is the technique of “alternating
wetness.” The time between dried and moist conditions does not seem to be long
enough to allow the soil to change from aerobic to anaerobic conditions, in contrast
to midseason drainage (Wassmann & Aulakh, 2000). Although alternating soaking
and drying minimized CH4 emissions, the N2O emissions from this system varied
widely. Methane can be oxidized and prevented by drainage and the resulting aero-
bic conditions of soil. According to Katayanagi et al. (2016), alternating paddy and
drying conditions can lessen CH4 emissions by 73% as compared to conventional
flooded rice. Correct watering is carried out according to the physiological param-
eters of the different growth phases of the plants. However, it can minimize the
frequency of alternating wet and dry environments, thereby reducing N2O and CH4
production and emissions. However, more research is needed to overcome the N2O
offset in this strategy.
Drainage Patterns
Intermittent drainage is defined as alternating periods of unrestricted drainage and
watering. It improves the oxidative conditions in the soil by improving root activity
and soil-bearing capacity and lowering anaerobic water input. It increases the aero-
bic surface area of the soil while reducing CH4 production by improving oxygen
transport. According to Yagi et al. (1996), sporadic drainage can reduce CH4 emis-
sions up to 44% as compared to conventional flooding. According to Adhya et al.
(2000), alternating drainage can reduce CH4 emissions by 15% compared to con-
tinuous flooding. In paddy fields, different water regimes lead to significant changes
in N2O emissions (Zou et al., 2005). However, Ye et al. (2013) found that intermit-
tent irrigation produced GWP (CH4 and N2O) 34% and 54% lower than flooding,
respectively.
Irrigation Control
Controlling irrigation has also been shown to lower net greenhouse gas emissions
from uprooted flooded rice (Hou et al., 2012). Similar to water management tech-
niques employed in rice-intensive systems, after replanting rice seedlings through-
out the rice production period in the absence of floods, the soil of flooded paddy
fields stays dry (60–80%) (Liu et al., 2015). Comparing paddy fields irrigated with
control irrigation with paddy fields irrigated with conventional flood irrigation,
Yang et al. (2014) found a 79% decline in CH4 emissions, a 10% increase in N2O
emissions, and a 67% reduction in GWP.
According to Hou et al. (2016), this irrigation system produced 27% less GWP
(CH4 and N2O) than conventional flooding. Additionally, contrary to irrigated rice,
Yan et al. (2021) and some writers testified that constant wet, water-saving irriga-
tion and deep irrigation (10-cm water depth) generate greenhouse gas emissions,
particularly CH4.
Cultivated Genotypes
Changes in soil characteristics, like soil temperature, porosity, soil moisture, etc.,
and biochemical activities have a significant effect on greenhouse gas emissions
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 17

from paddy fields due to tillage (Wu et al., 2013). By aerating the soil and physically
rupturing soil aggregates to liberate protected organic carbon components, soil dis-
turbance brought on by tillage can be intensified (Bilen et  al., 2022). Tillage
enhances soil carbon oxidation to carbon dioxide by enhancing soil aeration,
increasing crop residue-soil contact, and bare aggregate preservation of soil orgainc
matter (SOM) to microbial strike (Khaliq et al., 2013). Reducing tillage and soil
degradation in rice-dependent farming processes can reduce greenhouse gas emis-
sions. Guo and Zhou (2007) reported that no-till (NT) following the harvest of
spring wheat produced lower CO2 fluxes than regular tillage (CT). That is why NT
paddy fields generate a less amount of CH4 as compared to CT paddy fields due to
a number of factors (Li et al., 2013).
Compared to CT, Sakai et al. (2007) observed a 43% decrease in rice fields in
Japan’s Northern Territory’s seasonal cumulative CH4 emissions. Ghimire et  al.
(2017) found that with rice-wheat cropping systems, lowering the tillage frequency
dramatically reduced CH4 fluxes. The rise in soil bulk density, which results in a
lower volume of macropores and a reduction in the decay of humus, is what causes
the drop in CH4 emissions below LT. In a dry agricultural setting, Omonode et al.
(2007) discovered that NT-caused soil top hardening prevented CH4 from entering
the soil for oxidation, resulting in a reduction in CH4 absorption in the soil. Similarly,
increasing soil compaction may extend CH4 transport pathways, limiting CH4 emis-
sions into the atmosphere or through rice crops, or CH4 transfer to the rhizosphere
(Bassett et al., 2005).

5 Management of Organic Additives

In rice fields, organic additives have a significant impact on greenhouse gas emis-
sions. In general, adding the organic materials such as straw or organic fertilizers
leads to an increase in CH4 emissions, the extent of which varies depending on the
quantity, quality, and programing of application (Naser et al., 2007). In addition,
organic supplements produce freely available nitrogen pools in the soil, thereby
increasing N2O emissions (Tao et al., 2015). In fact, there is selected conflicting data
showing that high straw improvement reduces N2O emissions from paddy fields,
which can be attributed to N immobilization (Kreye et al., 2007). Understanding
this relationship presents complex challenges that need to be addressed depending
on the ecological situation.

5.1 Straw/Residue Management

Arable farming always generates large amounts of straw and other field waste. As
organic fertilizer use decreases, rice paddies are increasingly relying on straw recy-
cling to offset carbon losses from tillage and harvesting. Although straw burning
18 H. Bashir et al.

accelerates farmers’ seedbed preparation and reduces the hazard of nitrogen fixation
during the decomposition of residues with higher C/N ratios, incomplete C incin-
eration releases significant amounts of greenhouse gases and reduces air standards
(Hussain et al., 2015). In addition, nitrogen oxides and further burning of natural
chemicals contribute to the formation of tropospheric ozone. Rice straw contains a
variety of organic components, including cellulose, lipids, hemicellulose, lignin,
proteins, etc., and each component’s contribution to increasing the CH4 emission
rate varies. The CH4 discharge levels are most delicate for handling of straw in the
soil. Bugna et al. (1996) observed that off-season CH4 discharge rates from fresh
straw were higher than those from straw incorporated into rice fields. In a field study
conducted in Zhejiang Province, China, Deininger et al. (2000) reported that timely
straw bedding at the onset of winter fallow reduced GHG emissions by 11% in
comparison to traditional spring straw bedding methods.
Similarly, Wassmann et al. (2000b) demonstrated that remnant absorption under
the uncultivated phase (60 days before rice sowing) had advantages over traditional
pre-sowing treatment in terms of GHG emissions and crop production. In compari-
son to straw wrapping, removing straw reduced emissions of all three gases, sug-
gesting that eliminating the use of straw in paddy fields could also be a beneficial
approach. Koga and Tajima (2011) discovered that straw removal treatment had
lower CH4 and CO2 emissions than straw recirculation treatment. In the addition of
straw according to Bhattacharyya et al. (2013) CH4 emissions increased by 108%
and N2O emissions decreased up to 21% as related to manure plots. It also promotes
soil carbon sequestration, but its effect on CH4 growth is such that the GHG benefits
of reducing N2O emissions or increasing soil carbon sequestration are hardly out-
weighed. Because the decomposition of straw promotes the growth of methano-
gens, most of the CH4 is produced in flooded environments as straw decomposes. In
rice fields, Zschornack et  al. (2011) found that surface reservation of straw can
reduce CH4 and N2O emissions by 69% and 81%, respectively, in contrast to straw
integration. In rice-wheat cropping systems, the integration of wheat straw resulted
in more emissions than rice straw. From an economic and environmental point of
view, wheat straw mulch, whether partial or complete, is still beneficial. Besides
having a smaller effect on N2O emissions, mulching also significantly lowered CH4
emissions compared to capture (Ma et al., 2009).

5.2 Biochar Enrichment

Biochar is a carbon-rich by-product that results from the anaerobic pyrolysis of


waste biomass at high temperatures (Hussain et al., 2021a, 2022; Warnock et al.,
2007). The highly permeable structure of biochar, the finely grained C content, and
the enhanced surface make it a suitable soil additive for carbon sequestration
(Güereña et al., 2013). Use of biochar from pyrolysis of crop straw can raise carbon
sequestration up to 22% and lower GHG emissions up to 35% in relation to plots
without biochar. N2O emission from rice fields is more effective at higher biochar
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 19

concentrations (Shan et  al., 2013). Liu et  al. (2012) similarly documented large
reductions in GHG emissions through the use of biochar.

5.3 Selection of Rice Varieties

Choosing the right variety has been recognized as a viable way to reduce green-
house gas emissions, i.e., CH4 in rice soil. Rice cultivars (and their relatives) form
an extremely diverse crop (De Leon & Carpena, 1995), having greater than 90,000
attainments in the IRRI gene pool, each with unique morphophysiological traits and
environmental adaptations, and there are more than 90,000 accessions in the gene
pool (Price et al., 1999). In addition, the GWP of rice varieties varies greatly. Several
researches carried out in controlled and natural settings (Hussain et al. 2015) dem-
onstrated the diversity of rice cultivars and varieties in relation to the CH4 emission.
Differences in CH4 emissions between varieties are related to differences in CH4
liberation, oxidation, and transport capacity (Jiang et al., 2013). The threshold for
soil Eh is known to be 150 mV, which mainly controls the CH4 production rate in
rice soils (Yu & Patrick, 2004). Han et al. (2022) showed that soil Eh is regulated by
root respiration and exudation, aboveground biomass, and the growth state of the
rice plant during the rice planting season.
All of these factors are frequently listed as anticipated characteristics for the
development of CH4 budgets in paddy fields (Ding et al., 2004). The ability of rice
soils to generate and release CH4 is directly impacted by the dynamic fluctuations in
soil Eh and the changes in these parameters. Rice plants have variable CH4 oxidation
and differential O2 diffusion into the rhizosphere through the aerenchyma due to
changes in cultivar gas conductance connected to the release of O2 in the rhizo-
sphere (Huang et  al., 2021). Concentration gradients, diffusivity, internal aeren-
chyma structure, canopy density, root biomass, root pattern, overall biomass, and
metabolic activity all affect how quickly gases move through the tissue (Aulakh
et al., 2002). In order to sustain aerobic metabolism, a well-developed respiratory
tissue system makes sure that oxygen is available to the rhizosphere. This also limits
the possibility of potentially hazardous chemicals to enter plant roots through oxi-
dation (Armstrong et al., 1994). This promotes CH4 oxidation, which lowers atmo-
spheric emissions of the gas (Neue et al., 1997). Additionally, driven by concentration
and/or pressure gradients, it serves as a conduit for CH4 to reach the atmosphere
from the rhizosphere. Zheng et al. (2014) reported that super-rice CH4 emissions
were much lower than conventional rice and claimed that oxidation rather than pro-
duction was the main reason for the reduction in CH4 emissions. With more robust
root systems, rice varieties can deliver more oxygen to the soil, improve environ-
mental resilience, and promote yields. Zheng et al. (2014) also observed that the
root oxidation activity of super-rice before topping and at the topping stage was
much higher than that of standard rice seed.
The aerenchyma of rice plants, which supplies root exudates and/or dead root
cells as substrates for methanogens and methanotrophs, also controls the transfer of
20 H. Bashir et al.

CH4 and O2 (Kerdchoechuen, 2005). Different rice varieties produce different


amounts and types of root exudates, which are a breakdown of recently expelled
organic materials in rice roots (Aulakh et al., 2002). Crop transport is responsible
for the majority of CH4 emissions and Jiang et al. (2013) supported this theory and
suggested that rice variety selection has great potential for reducing CH4 emissions.
Cultivation practices can mask seasonal variations in species-specific CH4 emis-
sions (Barman & Mitra, 2019) and can be influenced by cultivation strategies.
Komatsu et al. (2009) reported that rice cultivars reaped at 3 months release less
CH4 than cultivars harvested at 4 months, indicating that a shorter growing season
is a clear factor for choosing low-emitting cultivars. A promising solution to lower
greenhouse gas emissions from rice soils appears to be choosing rice cultivars with
low CH4 emissions and increased resource utilization order. Prior to cultivar evalu-
ation, it is necessary to look at the mechanisms causing exudate and aerenchyma
impacts in the field.

6 Changing Planting Practices

6.1 Rice No-Till Technology

Direct rice (DSR), which can minimize greenhouse gas emissions and adapt to cli-
mate hazards, has been acknowledged as a promising alternative to traditional pud-
dle transplant rice (TPR), which is a substantial source of greenhouse gas (GHG)
emissions (Lip et al., 2014). Farooq et al. (2011) asserted that DSR is thought to be
a water-retaining method that can dramatically lower greenhouse gas emissions,
notably CH4 emissions, without lowering agricultural yields. They assert that the
primary causes of the decrease in CH4 generation and emissions in DSR compared
to TPR are less soil modification and shorter flooding durations. Wassmann et al.
(2004) found that midseason drainage of DSR fields can reduce CH4 emissions by
up to 50%. The reduction in CH4 fluxes, however, may be counterbalanced by an
increase in N2O emissions from dry DSRs due to changes in water status (Zheng
et al., 2014). Jiao et al. (2006) reported that N2O production at DSR increased as the
redox potential approached 250 mV.
To limit CH4 and N2O emissions, they recommend controlling the water supply
to keep the soil’s redox potential in the midrange (100–200 mV). Since the critical
soil redox potential for N2O formation has been determined to be 250 mV, this range
is sufficient to inhibit CH4 synthesis and promote the reduction of N2O to N2 (Jiao
et  al., 2006). The DSR system reduces N2O emissions, but because of its lower
GWP, it offers a more promising growth regime. Kumar et al. (2019) discovered that
DSR had a 53% lower average GWP than TPR for the three greenhouse gases (CO2,
CH4, and N2O). According to Ahmad et al. (2009) with the NT approach in paddy
fields, the GWP of DSR may be further decreased. Wang et al. (2017) also reported
that DSR cropping systems have the efficiency to significantly lower greenhouse
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 21

gas emissions from paddy fields. Due to lower GWP and high productivity, DSR
cultivation systems can result in reduced CH4 and N2O emissions per rice yield unit.
However, more extensive research, including simultaneous measurement of green-
house gases under the influence of water management, agriculture, fertilizers, and
more, is needed to recommend more appropriate DSR production options while
reducing environmental concerns.

7 Conclusions

Rising rice consumption and future population increase have sparked worries about
stabilizing greenhouse gas emissions in order to reduce predicted global climate
change. Here, we give a thorough evaluation of rice crop management techniques
effective in lowering greenhouse gas emissions in conjunction with recent data. We
are unable to cover all gases in each segment due to data constraints, but we have
looked into the viability and potential of several alternative practices based on the
GWP of greenhouse gases, especially CH4 and N2O. We discovered that crop admin-
istration practices can lessen the expected global climate change caused by rice
farming. For example, the reduction potentials of intermittent irrigation, midseason
drainage, and regulated irrigation systems for CH4 and N2O were 27–64%, 34–54%,
and 27–67%, respectively, compared to standard flood irrigation. The introduction
of NT and conservation tillage practices as an alternative to conventional farming is
effective for GHG reduction and C-smart farming because they reduce overall GHG
emissions. Rice fields can potentially reduce greenhouse gas emissions if straw is
managed through surface retention or mulching and biochar/compost is produced
on behalf of being burned or incorporated. The use of fermented organic fertilizers
is also a possible reduction strategy. Adjusting fertilization to the needs of plants,
precise positioning, replacing urea with ammonium sulfate, adding potassium fertil-
izer, and using nitrification inhibitors are effective measures to reduce greenhouse
gas emissions. In this context, the selection of varieties with lower CH4 emissions
and improved resource use efficiency also represents an important opportunity.
DSR also appears to be the most promising cultivation method and the most suitable
option for TPR in terms of GWP. It is anticipated that using all of these suggested
strategies to cut greenhouse gas emissions will either preserve rice yields or at the
very minimal enhance aid usage ability in the absence of lowering productivity. But
for these techniques to be effective, all social, economic, educational, and political
barriers must be eliminated. In order to build site-specific mitigation measures,
future research might concentrate on confirming the viability of these techniques in
various geographic locations and under other circumstances. Combining geographic
data, production and greenhouse gas emission models, and socioeconomic data
might also be beneficial. Finding cultivars with fewer greenhouse gas emissions
may be done using GIS data. Finding the GWP of various agricultural systems is
crucial. The GWP should ideally be determined using a standardized procedure.
There is a growing understanding that other aspects need to be taken into account
22 H. Bashir et al.

when discussing global climate change and agriculture, such as cultural signifi-
cance, supply of ecosystem assistance, food protection, and human health. We hope
that our efforts will be of use in the future, as their practical consequences will
inspire and guide future research. With regard to food security, any future proposals
to prevent catastrophic climate change should take this into account.

Acknowledgements  Authors thank the Higher Education Commission, Pakistan, for awarding
fellowship under International Research Support Initiative Program (IRSIP) Muhammad Mahroz
Hussain No. 1-8/HEC/HRD/2020/10831 and to the Environmental Biogeochemistry Laboratory,
University of Agriculture Faisalabad, Pakistan.  Author also want to thank UKRI-GCRF South
Asian Nitrogen Hub for providing the funding source. 

References

Adhya, T.  K., Bharati, K., Mohanty, S.  R., Ramakrishnan, B., Rao, V.  R., Sethunathan, N., &
Wassmann, R. (2000). Methane emission from rice fields at Cuttack, India. Nutrient Cycling in
Agroecosystems, 58(1/3), 95–105. https://doi.org/10.1023/A:1009886317629
Ahmad, S., Li, C., Dai, G., Zhan, M., Wang, J., Pan, S., & Cao, C. (2009). Greenhouse gas emis-
sion from direct seeding paddy field under different rice tillage systems in central China. Soil
and Tillage Research, 106(1), 54–61. https://doi.org/10.1016/j.still.2009.09.005
Armstrong, W., Brändle, R., & Jackson, M. B. (1994). Mechanisms of flood tolerance in plants.
Acta Botanica Neerlandica, 43(4), 307–358. https://doi.org/10.1111/j.1438-­8677.1994.
tb00756.x
Aulakh, M. S., Wassmann, R., & Rennenberg, H. (2002). Methane transport capacity of twenty-­
two rice cultivars from five major Asian rice-growing countries. Agriculture, Ecosystems and
Environment, 91(1–3), 59–71. https://doi.org/10.1016/S0167-­8809(01)00260-­2
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting circular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Barman, M., & Mitra, A. (2019). Temporal relationship between emitted and endogenous floral
scent volatiles in summer- and winter-blooming Jasminum species. Physiologia Plantarum,
166(4), 946–959. https://doi.org/10.1111/ppl.12849
Bassett, I. E., Simcock, R. C., & Mitchell, N. D. (2005). Consequences of soil compaction for
seedling establishment: Implications for natural regeneration and restoration. Austral Ecology,
30(8), 827–833. https://doi.org/10.1111/j.1442-­9993.2005.01525.x
Bhattacharyya, P., Nayak, A. K., Mohanty, S., Tripathi, R., Shahid, M., Kumar, A., Raja, R., Panda,
B. B., Roy, K. S., Neogi, S., Dash, P. K., Shukla, A. K., & Rao, K. S. (2013). Greenhouse gas
emission in relation to labile soil C, N pools and functional microbial diversity as influenced
by 39 years long-term fertilizer management in tropical rice. Soil and Tillage Research, 129,
93–105. https://doi.org/10.1016/j.still.2013.01.014
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 23

Bilen, S., Jacinthe, P.-A., Shrestha, R., Jagadamma, S., Nakajima, T., Kendall, J. R. A., Doohan, T.,
Lal, R., & Dick, W. (2022). Greenhouse gas fluxes in a no-tillage chronosequence in Central
Ohio. Soil and Tillage Research, 218, 105313. https://doi.org/10.1016/j.still.2021.105313
Bugna, G. C., Chanton, J. P., Cable, J. E., Burnett, W. C., & Cable, P. H. (1996). The importance of
groundwater discharge to the methane budgets of nearshore and continental shelf waters of the
northeastern Gulf of Mexico. Geochimica et Cosmochimica Acta, 60(23), 4735–4746. https://
doi.org/10.1016/S0016-­7037(96)00290-­6
Bundhoo, Z. M. A., Mauthoor, S., & Mohee, R. (2016). Potential of biogas production from bio-
mass and waste materials in the Small Island Developing State of Mauritius. Renewable and
Sustainable Energy Reviews, 56, 1087–1100. https://doi.org/10.1016/j.rser.2015.12.026
De Leon, J., & Carpena, A. (1995). Pedigree-based genetic diversity analysis of improved rice
(Oryza sativa L.) varieties in The Philippines. Philippine Journal of Crop Science (Philippines),
20, 1–12.
Deininger, A., Tamm, M., Krause, R., & Sonnenberg, H. (2000). SE—Structures and environment:
penetration resistance and water-holding capacity of differently conditioned straw for deep
litter housing systems. Journal of Agricultural Engineering Research, 77(3), 335–342. https://
doi.org/10.1006/jaer.2000.0606
Din, M. S. U., Mubeen, M., Hussain, S., Ahmad, A., Hussain, N., Ali, M. A., & Nasim, W. (2022).
World nations priorities on climate change and food security. In Building Climate Resilience in
Agriculture (365–384). Springer, Cham. https://doi.org/10.1007/978-3-030-79408-8_22
Ding, W., Cai, Z., & Wang, D. (2004). Preliminary budget of methane emissions from natural
wetlands in China. Atmospheric Environment, 38(5), 751–759. https://doi.org/10.1016/j.
atmosenv.2003.10.016
Epule, E.  T., Peng, C., & Mafany, N.  M. (2011). Methane emissions from paddy rice fields:
Strategies towards achieving a win-win sustainability scenario between rice production and
methane emission reduction. Journal of Sustainable Development, 4(6), 188. https://doi.
org/10.5539/jsd.v4n6p188
Farooq, M., Siddique, K. H. M., Rehman, H., Aziz, T., Lee, D.-J., & Wahid, A. (2011). Rice direct
seeding: Experiences, challenges and opportunities. Soil and Tillage Research, 111(2), 87–98.
https://doi.org/10.1016/j.still.2010.10.008
Flynn, H. C., & Smith, P. (2010). Greenhouse gas budgets of crop production: Current and likely
future trends. International Fertilizer Industry Association.
Ghimire, R., Lamichhane, S., Acharya, B.  S., Bista, P., & Sainju, U.  M. (2017). Tillage, crop
residue, and nutrient management effects on soil organic carbon in rice-based cropping sys-
tems: A review. Journal of Integrative Agriculture, 16(1), 1–15. https://doi.org/10.1016/
S2095-­3119(16)61337-­0
Güereña, D., Lehmann, J., Hanley, K., Enders, A., Hyland, C., & Riha, S. (2013). Nitrogen dynam-
ics following field application of biochar in a temperate North American maize-based produc-
tion system. Plant and Soil, 365(1–2), 239–254. https://doi.org/10.1007/s11104-­012-­1383-­4
Guo, J., & Zhou, C. (2007). Greenhouse gas emissions and mitigation measures in Chinese agroeco-
systems. Agricultural and Forest Meteorology, 142(2–4), 270–277. https://doi.org/10.1016/j.
agrformet.2006.03.029
Han, L., Chen, L., Li, D., Ji, Y., Feng, Y., Feng, Y., & Yang, Z. (2022). Influence of polyethyl-
ene terephthalate microplastic and biochar co-existence on paddy soil bacterial community
structure and greenhouse gas emission. Environmental Pollution, 292(B), 118386. https://doi.
org/10.1016/j.envpol.2021.118386
Hou, H., Peng, S., Xu, J., Yang, S., & Mao, Z. (2012). Seasonal variations of CH4 and N2O
emissions in response to water management of paddy fields located in Southeast China.
Chemosphere, 89(7), 884–892. https://doi.org/10.1016/j.chemosphere.2012.04.066
Hou, H., Yang, S., Wang, F., Li, D., & Xu, J. (2016). Controlled irrigation mitigates the annual
integrative global warming potential of methane and nitrous oxide from the rice–winter
wheat rotation systems in Southeast China. Ecological Engineering, 86, 239–246. https://doi.
org/10.1016/j.ecoleng.2015.11.022
24 H. Bashir et al.

Huang, C., Chen, X., Liu, L., Zhang, H., Yuan, B., & Li, Y. (2021). The influence of opening
shape of obstacles on explosion characteristics of premixed methane–air with concentra-
tion ­gradients. Process Safety and Environmental Protection, 150, 305–313. https://doi.
org/10.1016/j.psep.2021.04.028
Hussain, S., Peng, S., Fahad, S., Khaliq, A., Huang, J., Cui, K., & Nie, L. (2015). Rice man-
agement interventions to mitigate greenhouse gas emissions: A review. Environmental
Science and Pollution Research International, 22(5), 3342–3360. https://doi.org/10.1007/
s11356-­014-­3760-­4
Hussain, M., Farooqi, Z., Mohy-Ud-Din, W., Younas, F., Shahzad, M., Ghani, M., Ayub, M., &
Qadeer, A. (2021a). Application of bioremediation as sustainable approach to remediate heavy
metal and pesticide polluted environments. Plant and Environment, 1, 62–92.
Hussain, M. M., Bibi, I., Niazi, N. K., Shahid, M., Iqbal, J., Shakoor, M. B., Ahmad, A., Shah,
N. S., Bhattacharya, P., Mao, K., Bundschuh, J., Ok, Y. S., & Zhang, H. (2021b). Arsenic bio-
geochemical cycling in paddy soil-rice system: Interaction with various factors, amendments
and mineral nutrients. Science of the Total Environment, 773, 145040. https://doi.org/10.1016/j.
scitotenv.2021.145040
Hussain, M.  M., Zur Farooqi, R.  F., & Din, W.  M. U. (2021c). Role of microorganisms as cli-
mate engineers: Mitigation and adaptations to climate change. In J. A. Parray, S. A. Bandh,
& N.  Shameem (Eds.), Climate change and microbes: Impacts and vulnerability. 1. Apple
Academic Press.
Hussain, M. M., Mohy-Ud-Din, W., Younas, F., Niazi, N. K., Bibi, I., Yang, X., Rasheed, F., &
Biochar, F. Z. U. R. (2022). A game changer for sustainable agriculture. Journal of Sustainable
Agriculture, 143–157. Springer.
Jiang, Y., Wang, L.-l., Yan, X.-j., Tian, Y.-l., Deng, A.-x., & Zhang, W.-j. (2013). Super rice crop-
ping will enhance rice yield and reduce CH4 emission: A case study in Nanjing, China. Rice
Science, 20(6), 427–433. https://doi.org/10.1016/S1672-­6308(13)60157-­2
Jiao, Z., Hou, A., Shi, Y., Huang, G., Wang, Y., & Chen, X. (2006). Water management influenc-
ing methane and nitrous oxide emissions from rice field in relation to soil redox and micro-
bial community. Communications in Soil Science and Plant Analysis, 37(13–14), 1889–1903.
https://doi.org/10.1080/00103620600767124
Katayanagi, N., Fumoto, T., Hayano, M., Takata, Y., Kuwagata, T., Shirato, Y., Sawano, S., Kajiura,
M., Sudo, S., Ishigooka, Y., & Yagi, K. (2016). Development of a method for estimating total
CH4 emission from rice paddies in Japan using the DNDC-Rice model. Science of the Total
Environment, 547, 429–440. https://doi.org/10.1016/j.scitotenv.2015.12.149
Kerdchoechuen, O. (2005). Methane emission in four rice varieties as related to sugars and organic
acids of roots and root exudates and biomass yield. Agriculture, Ecosystems and Environment,
108(2), 155–163. https://doi.org/10.1016/j.agee.2005.01.004
Khaliq, A., Matloob, A., Ihsan, M. Z., Abbas, R. N., Aslam, Z., & Rasool, F. (2013). Supplementing
herbicides with manual weeding improves weed control efficiency, growth and yield of direct
seeded rice. International Journal of Agriculture and Biology, 15, 191–199.
Koga, N., & Tajima, R. (2011). Assessing energy efficiencies and greenhouse gas emissions
under bioethanol-oriented paddy rice production in northern Japan. Journal of Environmental
Management, 92(3), 967–973. https://doi.org/10.1016/j.jenvman.2010.11.008
Komatsu, G., Arzhannikov, S.  G., Gillespie, A.  R., Burke, R.  M., Miyamoto, H., & Baker,
V. R. (2009). Quaternary paleolake formation and cataclysmic flooding along the upper Yenisei
River. Geomorphology, 104(3–4), 143–164. https://doi.org/10.1016/j.geomorph.2008.08.009
Kreye, C., Dittert, K., Zheng, X., Zhang, X., Lin, S., Tao, H., & Sattelmacher, B. (2007). Fluxes of
methane and nitrous oxide in water-saving rice production in North China. Nutrient Cycling in
Agroecosystems, 77(3), 293–304. https://doi.org/10.1007/s10705-­006-­9068-­0
Kumar, A., Nayak, A.  K., Das, B.  S., Panigrahi, N., Dasgupta, P., Mohanty, S., Kumar, U.,
Panneerselvam, P., & Pathak, H. (2019). Effects of water deficit stress on agronomic and physi-
ological responses of rice and greenhouse gas emission from rice soil under elevated atmo-
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 25

spheric CO2. Science of the Total Environment, 650(2), 2032–2050. https://doi.org/10.1016/j.


scitotenv.2018.09.332
Le Mer, J., & Roger, P. (2001). Production, oxidation, emission and consumption of methane
by soils: A review. European Journal of Soil Biology, 37(1), 25–50. https://doi.org/10.1016/
S1164-­5563(01)01067-­6
Li, C., Zhang, Z., Guo, L., Cai, M., & Cao, C. (2013). Emissions of CH4 and CO2 from double
rice cropping systems under varying tillage and seeding methods. Atmospheric Environment,
80, 438–444. https://doi.org/10.1016/j.atmosenv.2013.08.027
Li, Y., Meng, Z., Xu, Y., Shi, Q., Ma, Y., Aung, M., Cheng, Y., & Zhu, W. (2021). Interactions
between anaerobic fungi and methanogens in the rumen and their biotechnological potential
in biogas production from lignocellulosic materials. Microorganisms, 9(1), 190. https://doi.
org/10.3390/microorganisms9010190
Linquist, B.  A., Adviento-Borbe, M.  A., Pittelkow, C.  M., van Kessel, C., & van Groenigen,
K.  J. (2012). Fertilizer management practices and greenhouse gas emissions from rice sys-
tems: A quantitative review and analysis. Field Crops Research, 135, 10–21. https://doi.
org/10.1016/j.fcr.2012.06.007
Lip, G. Y., Kongnakorn, T., Phatak, H., Kuznik, A., Lanitis, T., Liu, L. Z., Iloeje, U., Hernandez,
L., & Dorian, P. (2014). Cost-effectiveness of apixaban versus other new oral anticoagulants
for stroke prevention in atrial fibrillation. Clinical Therapeutics, 36(2), 192–210.e20. https://
doi.org/10.1016/j.clinthera.2013.12.011
Liu, X.-y., Qu, J.-j., Li, L.-q., Zhang, A.-f., Jufeng, Z., Zheng, J.-w., & Pan, G.-x. (2012). Can
biochar amendment be an ecological engineering technology to depress N2O emission in
rice paddies?—A cross site field experiment from South China. Ecological Engineering, 42,
168–173. https://doi.org/10.1016/j.ecoleng.2012.01.016
Liu, H., Hussain, S., Zheng, M., Peng, S., Huang, J., Cui, K., & Nie, L. (2015). Dry direct-seeded
rice as an alternative to transplanted-flooded rice in Central China. Agronomy for Sustainable
Development, 35(1), 285–294. https://doi.org/10.1007/s13593-­014-­0239-­0
Ma, J., Ma, E., Xu, H., Yagi, K., & Cai, Z. (2009). Wheat straw management affects CH4 and
N2O emissions from rice fields. Soil Biology and Biochemistry, 41(5), 1022–1028. https://doi.
org/10.1016/j.soilbio.2009.01.024
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In S. A. Bandh (Ed.),
Climate change: The social and scientific construct. Springer Nature Publishing. https://link.
springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Naser, H. M., Nagata, O., Tamura, S., & Hatano, R. (2007). Methane emissions from five paddy
fields with different amounts of rice straw application in Central Hokkaido, Japan. Soil Science
and Plant Nutrition, 53(1), 95–101. https://doi.org/10.1111/j.1747-­0765.2007.00105.x
Neue, H. U., Wassmann, R., Kludze, H. K., Bujun, W., & Lantin, R. S. (1997). Factors and pro-
cesses controlling methane emissions from rice fields. Nutrient Cycling in Agroecosystems,
49(1/3), 111–117. https://doi.org/10.1023/A:1009714526204
Omonode, R.  A., Vyn, T.  J., Smith, D.  R., Hegymegi, P., & Gál, A. (2007). Soil carbon diox-
ide and methane fluxes from long-term tillage systems in continuous corn and corn–soy-
bean rotations. Soil and Tillage Research, 95(1–2), 182–195. doi:https://doi.org/10.1016/j.
still.2006.12.004, 182.
Price, A., Steele, K., Townend, J., Gorham, J., Audebert, A., Jones, M., & Courtois, B. (1999).
Mapping root and shoot traits in rice: Experience in UK, IRRI and WARDA.  In Genetic
improvement of rice for water-limited environments (pp. 257–273). International Rice Research
Institute.
Sakai, S., Imachi, H., Sekiguchi, Y., Ohashi, A., Harada, H., & Kamagata, Y. (2007). Isolation of
key methanogens for global methane emission from rice paddy fields: A novel isolate affili-
ated with the clone cluster rice cluster I. Applied and Environmental Microbiology, 73(13),
4326–4331. https://doi.org/10.1128/AEM.03008-­06
26 H. Bashir et al.

Schils, R. L. M., Verhagen, A., Aarts, H. F. M., & Šebek, L. B. J. (2005). A farm level approach
to define successful mitigation strategies for GHG emissions from ruminant livestock
­systems. Nutrient Cycling in Agroecosystems, 71(2), 163–175. https://doi.org/10.1007/
s10705-­004-­2212-­9
Shan, Q., Wang, Y., Chen, K., Liang, Z., Li, J., Zhang, Y., Zhang, K., Liu, J., Voytas, D. F., Zheng, X.,
Zhang, Y., & Gao, C. (2013). Rapid and efficient gene modification in rice and Brachypodium
using TALENs. Molecular Plant, 6(4), 1365–1368. https://doi.org/10.1093/mp/sss162
Shiratori, Y., Watanabe, H., Furukawa, Y., Tsuruta, H., & Inubushi, K. (2007). Effectiveness
of a subsurface drainage system in poorly drained paddy fields on reduction of
methane emissions. Soil Science and Plant Nutrition, 53(4), 387–400. https://doi.
org/10.1111/j.1747-­0765.2007.00171.x
Smith, P., Fang, C., Dawson, J.  J. C., & Moncrieff, J.  B. (2008). Impact of global warm-
ing on soil organic carbon. Advances in Agronomy, 97, 1–43. https://doi.org/10.1016/
S0065-­2113(07)00001-­6
Tao, R., Liang, Y., Wakelin, S. A., & Chu, G. (2015). Supplementing chemical fertilizer with an
organic component increases soil biological function and quality. Applied Soil Ecology, 96,
42–51. https://doi.org/10.1016/j.apsoil.2015.07.009
Vaghefi, N., Shamsudin, M. N., Radam, A., & Rahim, K. A. (2016). Impact of climate change on food
security in Malaysia: Economic and policy adjustments for rice industry. Journal of Integrative
Environmental Sciences, 13(1), 19–35. https://doi.org/10.1080/1943815X.2015.1112292
van der Gon, H.  D. (2000). Changes in CH4 emission from rice fields from 1960 to 1990s: 1.
Impacts of modern rice technology. Global Biogeochemical Cycles, 14(1), 61–72. https://doi.
org/10.1029/1999GB900096
Van Nguyen, N., & Ferrero, A. (2006). Meeting the challenges of global rice production. Paddy
and Water Environment, 4(1), 1–9. https://doi.org/10.1007/s10333-­005-­0031-­5. Springer.
Wang, J., Zhang, X., Xiong, Z., Khalil, M. A. K., Zhao, X., Xie, Y., & Xing, G. (2012). Methane
emissions from a rice agroecosystem in South China: Effects of water regime, straw incorpora-
tion and nitrogen fertilizer. Nutrient Cycling in Agroecosystems, 93(1), 103–112. https://doi.
org/10.1007/s10705-­012-­9503-­3
Wang, W., Peng, S., Liu, H., Tao, Y., Huang, J., Cui, K., & Nie, L. (2017). The possibility of replac-
ing puddled transplanted flooded rice with dry seeded rice in Central China: A review. Field
Crops Research, 214, 310–320. https://doi.org/10.1016/j.fcr.2017.09.028
Warnock, D. D., Lehmann, J., Kuyper, T. W., & Rillig, M. C. (2007). Mycorrhizal responses to bio-
char in soil–concepts and mechanisms. Plant and Soil, 300(1–2), 9–20. https://doi.org/10.1007/
s11104-­007-­9391-­5
Wassmann, R., & Aulakh, M. S. (2000). The role of rice plants in regulating mechanisms of methane
missions. Biology and Fertility of Soils, 31(1), 20–29. https://doi.org/10.1007/s003740050619
Wassmann, R., Buendia, L. V., Lantin, R. S., Bueno, C. S., Lubigan, L. A., Umali, A., Nocon, N. N.,
Javellana, A. M., & Neue, H. U. (2000a). Mechanisms of crop management impact on meth-
ane emissions from rice fields in los Baños, Philippines. Nutrient Cycling in Agroecosystems,
58(1/3), 107–119. https://doi.org/10.1023/A:1009838401699
Wassmann, R., Lantin, R.  S., Neue, H.  U., Buendia, L.  V., Corton, T.  M., & Lu, Y. (2000b).
Characterization of methane emissions from rice fields in Asia. III.  Mitigation options and
future research needs. Nutrient Cycling in Agroecosystems, 58(1/3), 23–36. https://doi.org/1
0.1023/A:1009874014903
Wassmann, R., Hien, N.  X., Hoanh, C.  T., & Tuong, T.  P. (2004). Sea level rise affect-
ing the Vietnamese Mekong Delta: Water elevation in the flood season and
implications for rice production. Climatic Change, 66(1/2), 89–107. https://doi.org/10.1023/
B:CLIM.0000043144.69736.b7
Wu, G., Li, L., Ahmad, S., Chen, X., & Pan, X. (2013). A dynamic model for vulnerability assess-
ment of regional water resources in arid areas: A case study of Bayingolin, China. Water
Resources Management, 27(8), 3085–3101. https://doi.org/10.1007/s11269-­013-­0334-­z
2  Rice Production Technologies in Reducing Methane Gas Emissions for Sustainable… 27

Yagi, K., Tsuruta, H., Kanda, K., & Minami, K. (1996). Effect of water management on meth-
ane emission from a Japanese rice paddy field: Automated methane monitoring. Global
Biogeochemical Cycles, 10(2), 255–267. https://doi.org/10.1029/96GB00517
Yan, Q., Xu, Y., Chen, L., Cao, Z., Shao, Y., Xu, Y., Yu, Y., Fang, C., Zhu, Z., Feng, G., &
Chen, M. (2021). Irrigation with secondary municipal-treated wastewater: Potential effects,
­accumulation of typical antibiotics and grain quality responses in rice (Oryza sativa L.).
Journal of Hazardous Materials, 410, 124655. https://doi.org/10.1016/j.jhazmat.2020.124655
Yang, G., Chen, H., Wu, N., Tian, J., Peng, C., Zhu, Q., Zhu, D., He, Y., Zheng, Q., & Zhang,
C. (2014). Effects of soil warming, rainfall reduction and water table level on CH4 emis-
sions from the Zoige peatland in China. Soil Biology and Biochemistry, 78, 83–89. https://doi.
org/10.1016/j.soilbio.2014.07.013
Ye, Y., Liang, X., Chen, Y., Liu, J., Gu, J., Guo, R., & Li, L. (2013). Alternate wetting and drying
irrigation and controlled-release nitrogen fertilizer in late-season rice. Effects on dry matter
accumulation, yield, water and nitrogen use. Field Crops Research, 144, 212–224. https://doi.
org/10.1016/j.fcr.2012.12.003
Yu, K., & Patrick, W. H. (2004). Redox window with minimum global warming potential contri-
bution from rice soils. Soil Science Society of America Journal, 68(6), 2086–2091. https://doi.
org/10.2136/sssaj2004.2086
Zheng, H., Huang, H., Yao, L., Liu, J., He, H., & Tang, J. (2014). Impacts of rice varieties and man-
agement on yield-scaled greenhouse gas emissions from rice fields in China: A meta-analysis.
Biogeosciences, 11(13), 3685–3693. https://doi.org/10.5194/bg-­11-­3685-­2014
Zou, J., Huang, Y., Jiang, J., Zheng, X., & Sass, R.  L. (2005). A 3-year field measurement of
methane and nitrous oxide emissions from rice paddies in China: Effects of water regime, crop
residue, and fertilizer application. Global Biogeochemical Cycles, 19(2), n/a–n/a. https://doi.
org/10.1029/2004GB002401
Zschornack, T., Bayer, C., Zanatta, J. A., Vieira, F. C. B., & Anghinoni, I. (2011). Mitigation of
methane and nitrous oxide emissions from flood-irrigated rice by no incorporation of winter
crop residues into the soil. Revista Brasileira de Ciência do Solo, 35(2), 623–634. https://doi.
org/10.1590/S0100-­06832011000200031
Chapter 3
Manure Management to Reduce Methane
Emissions

Abdullah Kaviani Rad, Hassan Etesami, Angelika Astaikina,


and Rostislav Streletskii

Abstract  As a result of human activities, the production of greenhouse gases into


the atmosphere has increased in the past few hundred years, which has contributed
to the phenomenon of climate change. As the emissions of methane, a significant
greenhouse gas, have risen tremendously in recent years, especially from the agri-
cultural sector, numerous studies have been undertaken to find efficient methods for
reducing these emissions. A major focus of this chapter is on reducing the amount
of methane emitted from livestock production, particularly animal waste, and offers
biological alternatives that include the use of biofertilizers and biochar, the micro-
bial conversion of methane, genetic modification, and biogas production. It is also
suggested that precision agriculture (PA) policies and digital innovations, such as
the Internet of Things (IoT), unmanned aerial vehicle (UAV), and robotic systems,
be implemented to carefully monitor the application of manure to farms.
Furthermore, government economic plans, such as offering financial assistance to
farmers to minimize greenhouse gas emissions, which can be beneficial in lowering
methane emissions. Although several strategies have been recommended to decrease
greenhouse gas emissions, there has been no discernible shift in atmospheric meth-
ane concentrations; this would suggest that regulations such as PA on farms are not
being carried out effectively. Climate change and global warming will worsen over
the next few decades if no urgent action is taken to reduce greenhouse gas emissions.

A. K. Rad (*)
Department of Natural Resources and Environmental Engineering, College of Agriculture,
Shiraz University, Shiraz, Iran
H. Etesami
Department of Soil Science, University of Tehran, Tehran, Iran
A. Astaikina
Eurasian Center for Food Security, Lomonosov Moscow State University, Moscow, Russia
R. Streletskii
Soil Science Faculty, Lomonosov Moscow State University, Moscow, Russia

© The Author(s), under exclusive license to Springer Nature 29


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_3
30 A. K. Rad et al.

Keywords  Agriculture · Climate change · Global warming · Precision farming ·


Livestock · Biofertilizer · Organic fertilizers · Biogas

1 Introduction

In the past five thousand years, methane (CH4) emissions have risen because of the
growing population, rice cultivation in Asia, the subsequent need for more farm-
land, and the increased use of wood for cooking and heating (Li et  al., 2009).
Considering that CH4 is a greenhouse gas (GHG) emitted by both natural and
anthropogenic systems, the analysis of its emissions in various spatial and temporal
dimensions is of paramount importance (VanderZaag et  al., 2014; Yusuf et  al.,
2012). The CO2 concentration in the atmosphere has increased by 33% since 1750,
while the CH4 concentration has increased by 75%. The global warming potential
(GWP) of CH4 is 25 times that of CO2, and it contributes to approximately 20% of
the greenhouse effect (Dalal et al., 2008). Methane levels in the atmosphere began
to rise in 2007 following a period of nearly zero growth for seven years, and from
2014 to 2018, the global level of CH4 was approximately two times that in 2007
(Fletcher & Schaefer, 2019). Figure 3.1a illustrates the trend of growing methane
emissions from 1990 to 2019. In recent years, methane has been subject to extensive
research because of its disastrous effects on global warming and the chemistry of
the atmosphere (Yusuf et al., 2012). Agriculture, the energy sector, and the waste
management industry are the three most significant contributors to human resources
of GHGs (Fig. 3.1b). CH4 emissions are highest in the agricultural industry, energy,
and finally the waste management sector. Dalal et al. (2008) pinpointed intestinal
fermentation in ruminant animals (59%), fossil fuel power generation (24%), and

Stationary and mobile


Biomass (3%) sources (1%)
Agriculture (manure) ( 4%)
Enteric
Coal mining (6%) fermentation
(29%)
Global Methane Emissions
9000000
Other Ag sources (7%)
8000000

Methane
7000000

Emissions
6000000
(kt of CO2 equivalent)

Wastewater (9%)

by Source
5000000

4000000

3000000

Rice
2000000
cultivation
1000000 (10%)

0
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
2010

2012
2013
2014
2015
2016
2017
2018
2019
2011

Oil and Gas (20%)


Landfaills (11%)

(A) (B)

Fig. 3.1  Global statistics regarding methane emissions during 1990–2019 (a) and its sources (b).
(Adapted from IGSD (2016) and World Bank (2020))
3  Manure Management to Reduce Methane Emissions 31

landfills and domestic sewage (15%) as the major sources of CH4 in Australia. The
emissions of gases that contribute to global warming in the European Union are
largely attributable to agriculture (De Cara et al., 2005). At the same time, agricul-
tural production is impacted by GHGs. Methane and nitrogen oxide (N2O) emis-
sions were shown to be negatively associated with agricultural GDP in China
(Rehman et al., 2020).
As a major subsector of the agricultural industry, the livestock sector is a primary
source of the GHGs ammonia, CH4, and N2O (Hou et al., 2015). Ruminants release
CH4 through enteric fermentation (Ramin & Huhtanen, 2013). Anaerobic fermenta-
tion is used by the microbial communities in ruminants to decompose food.
Approximately 30% of all CH4 released into the atmosphere derives from livestock
such as cattle, sheep, and goats (Black et al., 2021). Animal manure is among the
key agricultural inputs that have been used to boost soil fertility by supplying N, P,
K, S, Ca, Mg, and Na, as well as other microelements including Fe, Mn, Cu, and Zn.
It also enhances organic matter in the soil and exchangeable cations, as well as soil
aggregate durability, soil physical properties, soil infiltration, moisture retention,
and resistance to cracking (Bayu et al., 2005). In spite of the fact that animal manure
contributes to soil fertility, it is estimated that cattle dung adds approximately 240
million metric tons of methane to the atmosphere annually (Tauseef et al., 2013).
Approximately 14.5% of anthropogenic greenhouse gas emissions worldwide are
attributed to livestock production, which represents 7.1 gigatons of CO2-equivalent
annually. Manufacturing and processing animal feed, as well as enteric fermenta-
tion of ruminants, are the two leading causes of environmental pollution, each con-
tributing 45% and 39%, respectively, to total emissions (FAO, 2022). It has been
estimated that approximately 40% of total human GHG emissions results from
ruminant enteric fermentation and animal dung, and this percentage is likely to
increase dramatically in the coming decades (Key & Tallard, 2012; Lassey, 2008).
As a result, it is critical to take steps to reduce the gases emitted by livestock (Black
et al., 2021). The reduction of livestock’s contribution to global CH4 emissions has
been addressed through a variety of biological and political approaches. For
instance, from an economic viewpoint, Hynes et al. (2009) reported that it would be
feasible to reduce GHG emissions if subsidies were presented to growers based on
the CH4 reduction program in Ireland. It is well demonstrated that the application of
biodigesters in crop fields is an effective method for lowering GHG emissions and
generating green energy (Flesch et al., 2011). In a study conducted by Jeffery et al.
(2016), the application of biochar in paddy and acidic soils was suggested as a tech-
nique to lower CH4 emissions. Methanotrophic bacteria bioconversion of CH4 into
microbial compounds has been proven to be a cost-effective and ecologically
acceptable method (Cantera et al., 2018). Pickering et al. (2015) examined the pos-
sibility of genetically modifying ruminants in order to reduce CH4. This gas is an
abundant and inexpensive carbon feedstock that has the potential to be bioconverted
into beneficial industrial products through the use of methanotrophs (Hwang et al.,
2018). Ruminant CH4 production is influenced by factors such as consumption
level, food type, and feed quality; therefore, modifying ruminant diets can be influ-
ential (Broucek, 2014; Haque, 2018).
32 A. K. Rad et al.

Despite the benefits of the aforementioned options, climate regulations, notably


those geared toward reducing CH4 emissions, do not appear to have significantly
changed the upward trend of global GHG emissions (Jackson et al., 2020). Methane
emissions from the livestock and fossil fuel sectors in the United States have been
reported by Miller et al. (2013) to be higher than those reported by the Environmental
Protection Agency (EPA) and the Emissions Database for Global Atmospheric
Research (EDGAR). According to the results calculated by Höglund-Isaksson et al.
(2020), approximately 7.7 Pg of CH4 is expected to be released into the atmosphere
between 2020 and 2050, and it is difficult to eliminate this gas using current techno-
logical solutions. Consequently, strategies to combat CH4 and the resulting global
warming are urgently needed in order to maintain agricultural sustainability and
food security (Reay et al., 2018). It has been determined that a prompt decrease in
CH4 significantly increases the probability of keeping global warming under 1.5 °C,
as determined by Collins et al. (2018). The objective of this chapter is to present an
overview of some practical methods to alleviate CH4 emissions from agronomic
systems within the framework of different mechanical and biological approaches.

2 Biological Approach

2.1 Alternative Fertilizers to Animal Manure

There are a number of steps that can be taken in order to reduce the amount of meth-
ane gas released into the atmosphere, such as the substitution of animal manure for
bio-fertilizers and other forms of organic fertilizers that have been processed or
combining animal manure with organic fertilizer. Additionally, it has been demon-
strated that mineral fertilizers, which are commonly used by farmers on agricultural
land such as paddies, may also affect the concentration of CH4 in agricultural soils
(Malyan et al., 2016). In rice paddies, nitrogen fertilizers applied in the form of urea
may significantly increase CH4 due to a reduction in redox potential and a change in
soil pH, which stimulate methanogenesis (Wang et al., 1993). According to Malyan
et al. (2016), the opposite effect was observed after fertilization with ammonium
sulfate, ammonium thiosulfate, and single superphosphate (SSP). The reduction of
CH4 emissions associated with the use of mineral fertilizers has been found to be
associated with a reduction in CH4 emissions as a result of the inhibition of metha-
nogens, which would affect the soil community structure (Bodelier 2011). Therefore,
using biofertilizers instead of animal fertilizers can reduce the negative effects of
using these fertilizers. In addition to being highly potent, biofertilizers have the
advantage of being environmentally friendly as compared to conventional inorganic
fertilizers (Rad et  al., 2022a). The use of other organic fertilizers (soil organic
amendments), such as compost, vermicompost, and biochar (a carbon-rich material
produced during the pyrolysis process), can also reduce the negative effects of ani-
mal manure (e.g., a reduction in CH4 emissions). For instance, Agnihotri et  al.
3  Manure Management to Reduce Methane Emissions 33

(1999) demonstrated that the composting of cow dung and leaves decreased CH4
fluxes on rice fields. In addition to reducing CH4, the application of biochar as a
fertilizer in rice paddies has been shown to significantly reduce GHG emissions
(Pandey et al., 2014; Zhang et al., 2010). Upon adding biochar to soil, the following
changes were observed: (1) the number of methanotrophic proteobacteria increased,
and (2) the ratio of methanogens to methanotrophs decreased. Molecular analysis,
on the other hand, demonstrated that the inhibition of methanogenic archaeal growth
was not associated with a decrease in CH4 concentration when biochar was present.
A comparison of the positive effects of biofertilizer and organic fertilizers is pre-
sented in Fig. 3.2.
Biofertilizers, such as diazotrophs (Azotobacter, Ochrobactrum anthropi, and
Azospirillum), plant growth-promoting rhizobacteria (PGPR), purple non-sulfur
bacteria (PNSB), cyanobacteria, and a water fern named Azolla, are widely used in
rice cultivation in Southeast Asia. A study conducted by Pingak et al. (2014) recog-
nized that inorganic fertilizer was combined with methanotrophic bacteria, diazo-
trophic bacteria, Ochrobactrum anthropi, Azotobacter, and Azospirillum to reduce
GHG emissions as well as increase rice yield and growth. By stimulating root
growth and hair growth, these bacterial isolates enhanced O2 diffusion in flooded
soils (Bhardwaj et al., 2014). A higher rate of plant growth was obtained by purple
non-sulfur bacteria (PNSB) under anoxic salt stress conditions than by methano-
gens (Kantha et al., 2015). In this context, Kantha et al. (2015) demonstrated that
PNSB Rhodopseudomonas palustris strains TN114, PP803, and TK103 were effec-
tive biofertilizers in paddy fields by increasing rice yield and reducing CH4 emis-
sions. Two types of rice fields were studied: organic paddy fields and saline paddy
fields. According to the findings of this study, purple non-sulfur bacteria were found

CH4 Flux

Organic
fertilizer
Inorganic Biofertilizer
fertilizers Biochar

Microbial Improving
communities microbial
degradation communities

Fig. 3.2 CH4 emissions and soil microbial communities are affected by biofertilizers and organic
fertilizers
34 A. K. Rad et al.

to increase grain yields over controls, and only strain TN114 was found to increase
rice yields in organic paddy fields. An earlier study showed that Rhodopseudomonas
palustris strains, particularly strain PP803, prevented salt from adversely impacting
rice seedlings and inhibited CO2 and CH4 emissions.
Azolla (aquatic pteridophyte) and the cyanobacteria Anabaena azollae
(Nostocaceae family) are important for reducing atmospheric carbon and fixing
nitrogen in paddy soil ecosystems (Ali et al., 2014). Moreover, a study by Prasanna
et al. (2002) indicated that active oxidation of CH4 by cyanobacteria and/or Azolla
microphylla resulted in a reduction in the emission of this GHG from rice paddies.
As part of this laboratory study, seven strains of cyanobacteria were evaluated for
their ability to reduce CH4 concentrations in rice soil samples. In the presence of the
Synechocystis sp., the CH4 concentration decreased by 10–20 times compared to the
control without the addition of cyanobacteria, demonstrating the maximum effec-
tiveness. Using silicate fertilizer in combination with urea and A. azollae in paddy
soil, Ali et al. (2014) reported a reduction in the maximum level of seasonal CH4
flux by 12%. The experiment evaluated the effectiveness of five complexes of soil
amendments: (1) urea + rice straw compost, (2) urea + rice straw compost + silicate
fertilizer, (3) urea + sesbania biomass + silicate fertilizer, (4) urea + azolla biomass
+ cyanobacterial mixture + silicate fertilizer, and (5) urea + cattle manure compost
+ silicate fertilizer. The A. azollae-treated plots showed the lowest CH4 flux of the
three treatments during both rice-growing seasons. Malyan et al. (2021) observed a
similar reduction in CH4 production when blue-green algae and Azolla were applied
as rice biofertilizers. As a result, Azolla and cyanobacteria can reduce CH4 flux in
two ways: (i) actively or directly oxidizing CH4 in hydromorphic soils and (ii)
increasing the soil redox potential (Eh) and, consequently, decreasing CH4 produc-
tion in rice soil. Additionally, the application of Azolla, Methylobacterium oryzae,
and blue-green algae can also reduce the emission of N2O into the atmosphere.

2.2 Biogas Production

A series of strategies have been devised under the Kyoto Protocol since 1997,
including one that is known as the Clean Development Mechanism (CDM), which
is intended to reduce emissions of GHGs (Yacob et  al., 2005). Solar, wind, and
biomass energy sources have emerged as sources of renewable energy (RE) that
have improved the sustainability and environmental quality of the energy market
(Ishikawa et al., 2021). In terms of heat and power generation, biomass energy pro-
duction is extremely useful and reduces GHGs (Ahmed et  al., 2017). As a clean
energy source, biogas is produced by anaerobic treatment of biomass feedstock
such as manure, sewage sludge, and food waste in a digester (Fig. 3.3) (Mathieu
Dumont et al., 2013; Li et al., 2019). In addition to containing 50–70% CH4 and
30–50% CO2, as well as a limited number of other gases, biogas has a heating value
of 21–24 MJ.m−3. The number of biogas plants in India and China has been approxi-
mately 4 million and 27 million since the 1970s, respectively. Generally, manure
3  Manure Management to Reduce Methane Emissions 35

Electricity
Biogas Biomethane
Animal manure,
Heat
Sewage sludge,
Weeds, Industrial
wastes,
Agricultural crops Livestock bedding
and residues
Digestate Fertilizer

Compost

Fig. 3.3 Applications of biogas production from biomass sources. (Adapted from Rabii
et al. (2019))

management systems are installed in rural areas and are fueled by livestock manure
(Bond & Templeton, 2011). By reducing global CH4, manure management systems
can significantly reduce GHG emissions (Martinez et  al., 2003; Xiaohong et  al.,
2011). Approximately 978–1776 kg of CO2 would be reduced annually if fermenta-
tion is applied in cow ranches based on a survey by Marañón et al. (2011). Moreover,
biogas can eliminate solid waste. According to research conducted by Baldé et al.
(2016), biogas generation was able to eliminate 62% of the volatile solid waste (VS)
input within two years. Kivaisi and Rubindamayugi (1996) estimated that Tanzania
produces 468,100 tons of organic matter annually, including coffee residue, sisal,
sugar, and grain wastes. Methane generation was calculated to be 400 m−3/ton for
sisal pulp, 400 m−3/ton for sisal industrial wastewater, and 650 m−3/ton for Robusta
coffee waste. The potential substitution of fossil fuels by biogas may lower annual
CO2 by approximately one million tons. By processing wastes with the anaerobic
digestion system, annual CH4 emissions can be decreased to roughly 189 million
m−3. Consequently, the production of biogas can contribute to the management of
water and soil pollution, the mitigation of GHGs, and the development of RE (Wang
& Calderon, 2012).
Different-sized digesters produce biogas, which can be utilized for the simulta-
neous generation of heat and power in boilers or as biofuel for diesel engines
(Börjesson & Berglund, 2006). In recent years, biogas processing has become cru-
cial for replacing natural gas and supplying fuel for CNG automobiles (Makaruk
et al., 2010; Subramanian et al., 2013). Adding organic matter to the soil after diges-
tion in the digester is beneficial as an organic fertilizer because, according to a study
carried out in Hangzhou (China) by Lu et  al. (2000), CH4 in paddy fields where
digested residues from a biogas plant were added to the soil was reduced. It was
between 10% and 16% less than farms that directly received pig manure. Landfills
and animal ranches can be regarded as biogas reactors that, in addition to meeting
the energy demands of on-site operations, are also capable of boosting the electric-
ity network (Karapidakis et al., 2010). Zamorano et al. (2007) evaluated a municipal
waste landfill in southern Spain and observed that it was capable of producing
36 A. K. Rad et al.

250–550 N m−3/h of CH4 at a concentration of 45%, which could be used to generate


nearly 4,500,000 kWh yearly. Biogas obtained from public garbage has high con-
centrations of aromatic hydrocarbons, siloxanes, and some halogenated hydrocar-
bons. In addition, the biogas produced from food waste is rich in sulfur-containing
compounds such as H2S and SO2 (Li et al., 2019).
Considering its commercial implications, biogas can potentially produce CH4
and N2O.  Although upgrading biogas to biomethane can improve air quality and
reduce GHGs, CH4 loss is still a significant environmental and economic challenge
for biogas (Paolini et al., 2018). Due to the fact that CH4 is 25 times more potent
than CO2, even minor biogas leaks have a substantial influence on global warming
(Mathieu Dumont et al., 2013). Recently, the environmental and economic effects
of CH4 loss from biogas power plants have been addressed (Kvist & Aryal, 2019).
Minimizing CH4 leaks from biogas sites is also crucial for maintaining efficiency
(Börjesson & Berglund, 2006). According to one estimate, CH4 losses from biogas
plants in the United Kingdom amounted to up to 3.8% of total emissions (Bakkaloglu
et al., 2021). Scheutz and Fredenslund (2019) assessed 23 biogas reactors and deter-
mined that CH4 loss rates varied from 0.4% to 14.9%. Meyer-Aurich et al. (2012)
revealed that GHG emissions resulting from the utilization of energy crop wastes or
CH4 leaks from reservoirs can neutralize the advantages of biogas.
The second challenge associated with the anaerobic digestion process is the
absence of appropriate optimization conditions. The effectiveness of biogas produc-
tion is influenced by variables such as feedstock selection, storage tanks, thermal
energy use, and ambient conditions (Meyer-Aurich et  al., 2012). According to
research conducted by Reinelt and Liebetrau (2020), monitoring a biogas plant
reveals that the rate of CH4 emission is dependent on a number of factors. Some of
these factors include shifts in the ambient temperature, the ability to use heat and
combined power, and the effectiveness of gasholders. In addition to operational
conditions such as tank filling level and substrate type, Hrad et al. (2015) demon-
strated that weather conditions such as wind speed and solar radiation also influence
CH4 flux. Muha et al. (2015) created a CH4 estimation model based on data from 21
biogas reactors in Germany and found that CH4 production efficiency is highly reli-
ant on factors such as hydraulic retention time (HRT), digestate removal rate from
the tank, and mixing ratio. Ruile et al. (2015) similarly documented a substantial
correlation between CH4 output and HRT. Biogas contains a variety of gases, includ-
ing H2S, which is hazardous and harms the pipelines and generators of an anaerobic
digestion system. Andriamanohiarisoamanana et al. (2018) observed that the addi-
tion of waste iron powder (WIP) at a rate of 2 g L−1 reduced H2S by 93%. In addi-
tion, it had no adverse influence on the anaerobic digestion process. According to
Farghali et al. (2020), direct mixing of WIP with cow dung is a feasible and cost-­
effective method for H2S removal and biogas purification since, in their investiga-
tion, adding 100 mg, 500 mg, and 1000 mg L−1 of WIP enhanced CH4 production
by 99.36%.
Presented herein are two methods that can be used in order to quantify the CH4
emissions from biogas plants: (i) an on-site measurement approach and (ii) a remote
3  Manure Management to Reduce Methane Emissions 37

sensing approach. As a result of a summation of the calculated rates between both


approaches, it is evident that there is a large difference between the results of both
approaches (Fredenslund et  al., 2018). As a result of using remote sensing tech-
niques to estimate CH4 losses at a power plant in Rhineland-Palatinate, Germany, it
was found that the average CH4 loss rate was 2.8 grams per second, which accounted
for 4% of the total biomass produced at the plant (Groth et al., 2015). As part of the
process of improving the biogas industry in the future, operational support networks
need to be strengthened (Bond & Templeton, 2011). There is no doubt that cities,
along with farms and industries, contribute significantly to GHG emissions (~70%),
but fewer efforts have been put into reducing CH4 flux from cities in comparison to
other sources. It has been concluded that in order to deal with CH4 in general, it is
necessary to invest in research and development efforts, implement new reduction
schemes, and conduct accurate and continuous monitoring of the amount of this gas
in the atmosphere (Hopkins et al., 2016).

3 Precision Agriculture Approach

There are a number of negative impacts associated with the uncontrolled use of
inputs in crop production that include waste of resources, destruction of the environ-
ment, as well as substantial financial losses for farmers (Bhattacharyay, 2020).
Taking appropriate action is necessary to ensure the long-term viability and optimal
performance of the food system. In order for the food system to remain viable, inno-
vative farming methods need to be adopted to combat climate change and food
insecurity (Demirbaş, 2018). It has been widely attributed to precision agriculture
(PA) as an advanced, all-encompassing, and globally standardized approach to man-
aging agricultural diversity and enhancing productivity (Schellberg et al., 2008). PA
contains novel agriculture management strategies that contribute to the aforemen-
tioned objectives. A great deal of investment has been made in agricultural research
and technology over the past few decades. This has made it apparent that PA is
crucial to the achievement of the goals of increased sustainable food production in
terms of yields, profits, and reduced environmental fallout. As a result of the new
and upgraded technologies that have enabled the collection of data, the food system
has become more transparent and safer (Fig.  3.4) (Demirbaş, 2018; Schellberg
et al., 2008). By focusing on the proper use of inputs, PA minimizes both the eco-
nomic costs associated with input preparation and the environmental damage asso-
ciated with agrochemical residues (Finger et  al., 2019; Rad et  al., 2022b). It is
possible to carry out a wide range of farming operations under the PA strategy, such
as soil preparation, planting, irrigation, weed control, spraying, and fertilizing. PA
aims to maximize the efficiency of these activities by using a variety of technologies
such as GPS, sensors, Internet of Things (IoT), robotics, unmanned aerial vehicles
(UAVs), machine learning, and decision support systems (Bhattacharyay, 2020;
Shamshiri et al., 2022). The environmental benefits of this approach include reduced
soil degradation, water contamination, carbon sequestration, and GHG emissions
38 A. K. Rad et al.

Fig. 3.4  Some of the


applicable technologies for
fertilizer management in IoT
agriculture. (Adapted from
Doshi and Varghese
(2022))
Machine Data
Learning Smart Analytics
Technologies
in
Agriculture

Cloud
Renewable Technology
Energy

(Kassam & Brammer, 2016; Malla et al., 2022; Bandh et al., 2021, 2023; Mushtaq
et al., 2020).
Precise fertilizer management is a useful framework for figuring out how PA
might be used to minimize GHG emissions, in particular CH4, from agricultural
systems. This recently developed concept integrates agronomic and manure man-
agement techniques with advanced technologies in order to improve crop yield
(Moshia et al., 2014). Fertilizer management includes the operation of collecting,
storing, and warehousing fertilizer, as well as the process of transporting fertilizer
to the field and applying it (Kleinman et al., 2017). Numerous studies have shown
that proper manure management can have positive effects on the economy and the
ecosystem (Niles et al., 2019). Increased profitability, a lower risk of soil pollution
from livestock manure, and enhanced crop and livestock yields are all major conse-
quences of efficient management of manure distribution on farms. Morris et  al.
(1999) developed a GPS-enabled fertilizer spreader with the purpose of avoiding
fertilization near water sources and other sensitive areas. By analyzing a fertilizer
distributor, Cabot et al. (2006) demonstrated that it is possible to precisely regulate
soil nitrogen levels using the instrument. Additionally, remote control technologies
such as satellite imagery, UAVs, and IoT have become serviceable and can dramati-
cally improve the above processes in farming (Higgins et  al., 2019). Using data
from three farms on the Portuguese-Spanish border, Loures et al. (2020) determined
that integrating systems such as remotely piloted aircraft systems (RPAS), UAV, and
Normalized Vegetation Difference Index (NDVI) can result in significant economic
savings, even on small farms (less than 50 ha). Evidently, the effectiveness of PA is
contingent on its precise application in analyzing the circumstances and spatial and
temporal administration of agriculture activities (Pierce & Nowak, 1999).
However, in spite of the growing number of advantages that PA technologies can
offer as well as the fact that PA technologies can be widely applied in fields, growers
3  Manure Management to Reduce Methane Emissions 39

still do not employ them (Lindblom et al., 2017). It has been reported that the rate
of adoption of PA by farmers at present is extremely low (Higgins et al., 2019), and
the reasons for the disparate rates of adoption between and within countries merit
further research (Kassam & Brammer, 2016). As Schieffer and Dillon (2013) found
in their survey of farmers in western Kentucky, the extra costs associated with
implementing PA technologies on farms make them less responsive to policies
designed solely to provide financial incentives to reduce pollution. The reluctance
to perform PA appears to be caused by a lack of access to specialists, as well as
financial and educational limitations (Kitchen et al., 2002). Governments, however,
may also contribute to the challenges associated with PA implementation. Kulyasov
et al. (2020) addressed the challenges associated with expanding PA in Russia. They
concluded that the absence of a regulatory framework for data management and the
lack of institutional support for PA were major factors. Stuart et  al. (2014) con-
ducted a survey of US corn farmers and found that poor perceptions of climate
change, lack of access to advanced technologies, and political and financial restric-
tions can all prevent the enhancement of nitrogen efficiency. By utilizing educa-
tional programs and tax incentives, some of these challenges can be addressed. PA
offers economic, social, and ecological advantages that can be enhanced through the
enhancement of its technical infrastructure and the adoption of legal frameworks
(Finger et al., 2019). As PA is essential for food security, it is imperative to enhance
(i) agricultural knowledge and (ii) computer science and information management
abilities in order to facilitate its development. Furthermore, a multidisciplinary
approach and the collaboration of professionals from other scientific disciplines are
required for agriculture’s long-term sustainability (Lindblom et al., 2017). To com-
bat the challenge of global warming and maintain agricultural production, interna-
tional collaboration is required (Rad et  al., 2022c; Zarei & Kaviani Rad, 2020;
Parray et al., 2022 Bandh et al., 2022 Bandh, 2022a, b).

4 Conclusions

Methane (CH4) is a greenhouse gas that is released into the atmosphere as a result
of natural and anthropogenic activities. These include deforestation and biomass
burning, rice farming, and livestock production. Due to intestinal fermentation in
ruminants and the production of manure, the livestock industry is widely regarded
as one of the largest contributors to CH4 emissions. In this chapter, practical meth-
ods for reducing the emissions of GHGs from livestock manure were examined. The
first efficient biological treatment option was the substitution or combined applica-
tion of animal dung and biofertilizers such as some PGPRs, fungi, and Azolla.
Additionally, biogas reactors can be advantageous for limiting CH4 produced from
animal manure if appropriate manufacturing conditions are provided and methane
loss is controlled. Measurement and management of manure distribution in the field
may be accomplished by utilizing technology such as the IoT, robotics, and UAVs.
To prevent further increases in CH4 production in the coming years and the
40 A. K. Rad et al.

subsequent worsening of global warming and climate change, the recommended


solutions must be implemented immediately on a global level.

References

Agnihotri, S., Kulshreshtha, K., & Singh, S.  N. (1999). Mitigation strategy to contain methane
emission from rice-fields. Environmental Monitoring and Assessment, 58(1), 95–104. https://
doi.org/10.1023/A:1006081317688
Ahmed, W. A., Aggour, M., & Naciri, M. (2017). Biogas control: Methane production monitoring
using Arduino. International Journal of Biotechnology and Bioengineering, 11(2), 130–133.
https://doi.org/10.5281/zenodo.1339948
Ali, M. A., Sattar, M. A., Islam, M. N., & Inubushi, K. (2014). Integrated effects of organic, inor-
ganic and biological amendments on methane emission, soil quality and rice productivity in
irrigated paddy ecosystem of Bangladesh: Field study of two consecutive rice growing seasons.
Plant and Soil, 378(1–2), 239–252. https://doi.org/10.1007/s11104-­014-­2023-­y
Andriamanohiarisoamanana, F.  J., Shirai, T., Yamashiro, T., Yasui, S., Iwasaki, M., Ihara, I.,
Nishida, T., Tangtaweewipat, S., & Umetsu, K. (2018). Valorizing waste iron powder in bio-
gas production: Hydrogen sulfide control and process performances. Journal of Environmental
Management, 208, 134–141. https://doi.org/10.1016/j.jenvman.2017.12.012
Bakkaloglu, S., Lowry, D., Fisher, R. E., France, J. L., Brunner, D., Chen, H., & Nisbet, E. G. (2021).
Quantification of methane emissions from UK biogas plants. Waste Management, 124, 82–93.
https://doi.org/10.1016/j.wasman.2021.01.011
Baldé, H., VanderZaag, A. C., Burtt, S. D., Wagner-Riddle, C., Crolla, A., Desjardins, R. L., &
MacDonald, D. J. (2016). Methane emissions from digestate at an agricultural biogas plant.
Bioresource Technology, 216, 914–922. https://doi.org/10.1016/j.biortech.2016.06.031
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting circular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Bayu, W., Rethman, N. F. G., & Hammes, P. S. (2005). The role of animal manure in sustainable
soil fertility management in sub-Saharan Africa: A review. Journal of Sustainable Agriculture,
25(2), 113–136. https://doi.org/10.1300/J064v25n02_09
Bhardwaj, D., Ansari, M. W., Sahoo, R. K., & Tuteja, N. (2014). Biofertilizers function as key
player in sustainable agriculture by improving soil fertility, plant tolerance and crop productiv-
ity. Microbial Cell Factories, 13(1), 66. https://doi.org/10.1186/1475-­2859-­13-­66
Bhattacharyay, D. (2020). Future of precision agriculture in India. In
Protected cultivation and smart agriculture. https://www.researchgate.net/
publication/347511698_Future_of_Precision_Agriculture_in_India
3  Manure Management to Reduce Methane Emissions 41

Black, J. L., Davison, T. M., & Box, I. (2021). Methane emissions from ruminants in Australia:
Mitigation potential and applicability of mitigation strategies. Animals (Basel), 11(4). https://
doi.org/10.3390/ani11040951
Bodelier, P. L. E. (2011). Interactions between nitrogenous fertilizers and methane cycling in wet-
land and upland soils. Current Opinion in Environmental Sustainability, 3(5), 379–388. https://
doi.org/10.1016/j.cosust.2011.06.002
Bond, T., & Templeton, M. R. (2011). History and future of domestic biogas plants in the devel-
oping world. Energy for Sustainable Development, 15(4), 347–354. https://doi.org/10.1016/j.
esd.2011.09.003
Börjesson, P., & Berglund, M. (2006). Environmental systems analysis of biogas systems—Part
I: Fuel-cycle emissions. Biomass and Bioenergy, 30(5), 469–485. https://doi.org/10.1016/j.
biombioe.2005.11.014
Broucek, J. (2014). Production of methane emissions from ruminant husbandry: A review. Journal
of Environmental Protection, 05(15), 1482–1493. https://doi.org/10.4236/jep.2014.515141
Cabot, P., Pierce, F., Nowak, P., & Karthikeyan, K. (2006). Monitoring and predicting manure appli-
cation rates using precision conservation technology. Journal of Soil and Water Conservation,
61(5), 282–292. https://www.jswconline.org/content/61/5/282
Cantera, S., Muñoz, R., Lebrero, R., López, J. C., Rodríguez, Y., & García-Encina, P. A. (2018).
Technologies for the bioconversion of methane into more valuable products. Current Opinion
in Biotechnology, 50, 128–135. https://doi.org/10.1016/j.copbio.2017.12.021
Collins, W. J., Webber, C. P., Cox, P. M., Huntingford, C., Lowe, J., Sitch, S., Chadburn, S. E.,
Comyn-Platt, E., Harper, A.  B., Hayman, G., & Powell, T. (2018). Increased importance of
methane reduction for a 1.5 degree target. Environmental Research Letters, 13(5). https://doi.
org/10.1088/1748-­9326/aab89c
Dalal, R.  C., Allen, D.  E., Livesley, S.  J., & Richards, G. (2008). Magnitude and biophysical
regulators of methane emission and consumption in the Australian agricultural, forest, and
submerged landscapes: A review. Plant and Soil, 309(1–2), 43–76. https://doi.org/10.1007/
s11104-­007-­9446-­7
De Cara, S., Houzé, M., & Jayet, P.-A. (2005). Methane and nitrous oxide emissions from agri-
culture in the EU: A spatial assessment of sources and abatement costs. Environmental and
Resource Economics, 32(4), 551–583. https://doi.org/10.1007/s10640-­005-­0071-­8
Demirbaş, N. (2018). Precision agriculture in terms of food security: Needs for the future. Precision
Agriculture, 27. https://www.researchgate.net/publication/328655146_Precision_Agriculture_
in_Terms_of_Food_Security_Needs_for_The_Future
Doshi, M., & Varghese, A. (2022). Chapter 12. Smart agriculture using renewable energy and
AI-powered IoT.  In A.  Abraham, S.  Dash, J.  J. P.  C. Rodrigues, B.  Acharya, & S.  K. Pani
(Eds.), AI, edge and IoT-based smart agriculture (pp. 205–225). Academic Press. https://doi.
org/10.1016/B978-­0-­12-­823694-­9.00028-­1
Farghali, M., Andriamanohiarisoamanana, F. J., Ahmed, M. M., Kotb, S., Yamamoto, Y., Iwasaki,
M., Yamashiro, T., & Umetsu, K. (2020). Prospects for biogas production and H2S control
from the anaerobic digestion of cattle manure: The influence of microscale waste iron pow-
der and iron oxide nanoparticles. Waste Management, 101, 141–149. https://doi.org/10.1016/j.
wasman.2019.10.003
Finger, R., Swinton, S. M., El Benni, N., & Walter, A. (2019). Precision farming at the nexus of
agricultural production and the environment. Annual Review of Resource Economics, 11(1),
313–335. https://doi.org/10.1146/annurev-­resource-­100518-­093929
Flesch, T. K., Desjardins, R. L., & Worth, D. (2011). Fugitive methane emissions from an agri-
cultural biodigester. Biomass and Bioenergy, 35(9), 3927–3935. https://doi.org/10.1016/j.
biombioe.2011.06.009
Fletcher, S.  E. M., & Schaefer, H. (2019). Rising methane: A new climate challenge. Science,
364(6444), 932–933. https://doi.org/10.1126/science.aax1828
Food and Agriculture Organization. (2022). Key facts and findings, GHG emissions by livestock.
https://www.fao.org/news/story/en/item/197623/icode/. Retrieved July 20, 2022.
42 A. K. Rad et al.

Fredenslund, A. M., Hinge, J., Holmgren, M. A., Rasmussen, S. G., & Scheutz, C. (2018). On-site
and ground-based remote sensing measurements of methane emissions from four biogas
plants: A comparison study. Bioresource Technology, 270, 88–95. https://doi.org/10.1016/j.
biortech.2018.08.080
Groth, A., Maurer, C., Reiser, M., & Kranert, M. (2015). Determination of methane emission rates
on a biogas plant using data from laser absorption spectrometry. Bioresource Technology, 178,
359–361. https://doi.org/10.1016/j.biortech.2014.09.112
Haque, M.  N. (2018). Dietary manipulation: A sustainable way to mitigate methane emissions
from ruminants. Journal of Animal Science and Technology, 60, 15. https://doi.org/10.1186/
s40781-­018-­0175-­7
Higgins, S., Schellberg, J., & Bailey, J. S. (2019). Improving productivity and increasing the effi-
ciency of soil nutrient management on grassland farms in the UK and Ireland using precision
agriculture technology. European Journal of Agronomy, 106, 67–74. https://doi.org/10.1016/j.
eja.2019.04.001
Höglund-Isaksson, L., Gómez-Sanabria, A., Klimont, Z., Rafaj, P., & Schöpp, W. (2020). Technical
potentials and costs for reducing global anthropogenic methane emissions in the 2050 time-
frame  – Results from the GAINS model. Environmental Research Communications, 2(2).
https://doi.org/10.1088/2515-­7620/ab7457
Hopkins, F. M., Ehleringer, J. R., Bush, S. E., Duren, R. M., Miller, C. E., Lai, C.-T., Hsu, Y.-K.,
Carranza, V., & Randerson, J.  T. (2016). Mitigation of methane emissions in cities: How
new measurements and partnerships can contribute to emissions reduction strategies. Earth’s
Future, 4(9), 408–425. https://doi.org/10.1002/2016EF000381
Hou, Y., Velthof, G. L., & Oenema, O. (2015). Mitigation of ammonia, nitrous oxide and meth-
ane emissions from manure management chains: A meta-analysis and integrated assessment.
Global Change Biology, 21(3), 1293–1312. https://doi.org/10.1111/gcb.12767
Hrad, M., Piringer, M., & Huber-Humer, M. (2015). Determining methane emissions from bio-
gas plants—Operational and meteorological aspects. Bioresource Technology, 191, 234–243.
https://doi.org/10.1016/j.biortech.2015.05.016
Hwang, I.  Y., Nguyen, A.  D., Nguyen, T.  T., Nguyen, L.  T., Lee, O.  K., & Lee, E.  Y. (2018).
Biological conversion of methane to chemicals and fuels: Technical challenges and issues.
Applied Microbiology and Biotechnology, 102(7), 3071–3080. https://doi.org/10.1007/
s00253-­018-­8842-­7
Hynes, S., Morrissey, K., O’Donoghue, C., & Clarke, G. (2009). A spatial micro-simulation analy-
sis of methane emissions from Irish agriculture. Ecological Complexity, 6(2), 135–146. https://
doi.org/10.1016/j.ecocom.2008.10.014
IGSD. (2016). Estimated global anthropogenic methane emissions by source, Retrieved July
17, 2022. https://www.igsd.org/estimated-­global-­anthropogenic-­methane-­emissions-­by-­
source-­2010/
Ishikawa, S., Connell, N.  O., Lechner, R., Hara, R., Kita, H., & Brautsch, M. (2021). Load
response of biogas CHP systems in a power grid. Renewable Energy, 170, 12–26. https://doi.
org/10.1016/j.renene.2021.01.120
Jackson, R. B., Saunois, M., Bousquet, P., Canadell, J. G., Poulter, B., Stavert, A. R., Bergamaschi,
P., Niwa, Y., Segers, A., & Tsuruta, A. (2020). Increasing anthropogenic methane emissions
arise equally from agricultural and fossil fuel sources. Environmental Research Letters, 15(7).
https://doi.org/10.1088/1748-­9326/ab9ed2
Jeffery, S., Verheijen, F. G. A., Kammann, C., & Abalos, D. (2016). Biochar effects on methane
emissions from soils: A meta-analysis. Soil Biology and Biochemistry, 101, 251–258. https://
doi.org/10.1016/j.soilbio.2016.07.021
Kantha, T., Kantachote, D., & Klongdee, N. (2015). Potential of biofertilizers from selected
Rhodopseudomonas palustris strains to assist rice (Oryza sativa L. subsp. indica) growth under
salt stress and to reduce greenhouse gas emissions. Annals of Microbiology, 65(4), 2109–2118.
https://doi.org/10.1007/s13213-­015-­1049-­6
3  Manure Management to Reduce Methane Emissions 43

Karapidakis, E.  S., Tsave, A.  A., Soupios, P.  M., & Katsigiannis, Y.  A. (2010). Energy effi-
ciency and environmental impact of biogas utilization in landfills. International Journal of
Environmental Science and Technology, 7(3), 599–608. https://doi.org/10.1007/BF03326169
Kassam, A., & Brammer, H. (2016). Environmental implications of three modern agricul-
tural practices: Conservation Agriculture, the System of Rice Intensification and Precision
Agriculture. International Journal of Environmental Studies, 73(5), 702–718. https://doi.org/1
0.1080/00207233.2016.1185329
Key, N., & Tallard, G. (2012). Mitigating methane emissions from livestock: A global analysis of sec-
toral policies. Climatic Change, 112(2), 387–414. https://doi.org/10.1007/s10584-­011-­0206-­6
Kitchen, N.  R., Snyder, C.  J., Franzen, D.  W., & Wiebold, W.  J. (2002). Educational needs
of precision agriculture. Precision Agriculture, 3(4), 341–351. https://doi.org/10.102
3/A:1021588721188
Kivaisi, A. K., & Rubindamayugi, M. S. T. (1996). The potential of agro-industrial residues for
production of biogas and electricity in Tanzania. Renewable Energy, 9(1–4), 917–921. https://
doi.org/10.1016/0960-­1481(96)88429-­1
Kleinman, P.  J. A., Buda, A.  R., Sharpley, A.  N., Khosla, R., Delgado, J., Sassenrath, G., &
Mueller, T. (2017). Elements of precision manure management. In Precision conservation
(pp. 165–192). https://doi.org/10.2134/agronmonogr59.c9.
Kulyasov, N. S., Novik, N. N., Klyukin, N. D., & Charyyarova, G. D. (2020). Precision agriculture in
The Russian Federation: Problems and directions in development. IOP Conference Series: Earth
and Environmental Science, 548(2), 022090. https://doi.org/10.1088/1755-­1315/548/2/022090
Kvist, T., & Aryal, N. (2019). Methane loss from commercially operating biogas upgrading plants.
Waste Management, 87, 295–300. https://doi.org/10.1016/j.wasman.2019.02.023
Lassey, K. R. (2008). Livestock methane emission and its perspective in the global methane cycle.
Australian Journal of Experimental Agriculture, 48(2). https://doi.org/10.1071/EA07220
Li, X., Dodson, J., Zhou, J., & Zhou, X. (2009). Increases of population and expansion of rice agri-
culture in Asia, and anthropogenic methane emissions since 5000 BP. Quaternary International,
202(1–2), 41–50. https://doi.org/10.1016/j.quaint.2008.02.009
Li, Y., Alaimo, C. P., Kim, M., Kado, N. Y., Peppers, J., Xue, J., Wan, C., Green, P. G., Zhang, R.,
Jenkins, B. M., Vogel, C. F. A., Wuertz, S., Young, T. M., & Kleeman, M. J. (2019). Composition
and toxicity of biogas produced from different feedstocks in California. Environmental Science
and Technology, 53(19), 11569–11579. https://doi.org/10.1021/acs.est.9b03003
Lindblom, J., Lundström, C., Ljung, M., & Jonsson, A. (2017). Promoting sustainable intensifica-
tion in precision agriculture: Review of decision support systems development and strategies.
Precision Agriculture, 18(3), 309–331. https://doi.org/10.1007/s11119-­016-­9491-­4
Loures, L., Chamizo, A., Ferreira, P., Loures, A., Castanho, R., & Panagopoulos, T. (2020).
Assessing the effectiveness of precision agriculture management systems in Mediterranean
small farms. Sustainability, 12(9). https://doi.org/10.3390/su12093765
Lu, W.  F., Chen, W., Duan, B.  W., Guo, W.  M., Lu, Y., Lantin, R.  S., Wassmann, R., & Neue,
H. U. (2000). Methane emissions and mitigation options in irrigated rice fields in Southeast
China. Nutrient Cycling in Agroecosystems, 58(1/3), 65–73. https://doi.org/10.102
3/A:1009830232650
Makaruk, A., Miltner, M., & Harasek, M. (2010). Membrane biogas upgrading processes for the
production of natural gas substitute. Separation and Purification Technology, 74(1), 83–92.
https://doi.org/10.1016/j.seppur.2010.05.010
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In S. A. Bandh (Ed.),
Climate change: The social and scientific construct. Springer Nature Publishing. https://link.
springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Malyan, S. K., Bhatia, A., Kumar, A., Gupta, D. K., Singh, R., Kumar, S. S., Tomer, R., Kumar, O.,
& Jain, N. (2016). Methane production, oxidation and mitigation: A mechanistic understanding
and comprehensive evaluation of influencing factors. Science of the Total Environment, 572,
874–896. https://doi.org/10.1016/j.scitotenv.2016.07.182
44 A. K. Rad et al.

Malyan, S. K., Bhatia, A., Tomer, R., Harit, R. C., Jain, N., Bhowmik, A., & Kaushik, R. (2021).
Mitigation of yield-scaled greenhouse gas emissions from irrigated rice through Azolla, Blue-­
green algae, and plant growth–promoting bacteria. Environmental Science and Pollution
Research International, 28(37), 51425–51439. https://doi.org/10.1007/s11356-­021-­14210-­z
Marañón, E., Salter, A. M., Castrillón, L., Heaven, S., & Fernández-Nava, Y. (2011). Reducing the
environmental impact of methane emissions from dairy farms by anaerobic digestion of cattle
waste. Waste Management, 31(8), 1745–1751. https://doi.org/10.1016/j.wasman.2011.03.015
Martinez, J., Guiziou, F., Peu, P., & Gueutier, V. (2003). Influence of treatment techniques for
pig slurry on methane emissions during subsequent storage. Biosystems Engineering, 85(3),
347–354. https://doi.org/10.1016/S1537-­5110(03)00067-­9
Mathieu Dumont, N. L., Luning, L., Yildiz, I., & Koop, K. (2013). Methane emissions in biogas pro-
duction. In The biogas handbook (pp. 248–266). https://doi.org/10.1533/9780857097415.2.248
.
Meyer-Aurich, A., Schattauer, A., Hellebrand, H.  J., Klauss, H., Plöchl, M., & Berg,
W. (2012). Impact of uncertainties on greenhouse gas mitigation potential of biogas produc-
tion from agricultural resources. Renewable Energy, 37(1), 277–284. https://doi.org/10.1016/j.
renene.2011.06.030
Miller, S.  M., Wofsy, S.  C., Michalak, A.  M., Kort, E.  A., Andrews, A.  E., Biraud, S.  C.,
Dlugokencky, E.  J., Eluszkiewicz, J., Fischer, M.  L., Janssens-Maenhout, G., Miller, B.  R.,
Miller, J. B., Montzka, S. A., Nehrkorn, T., & Sweeney, C. (2013). Anthropogenic emissions of
methane in the United States. Proceedings of the National Academy of Sciences of the United
States of America, 110(50), 20018–20022. https://doi.org/10.1073/pnas.1314392110
Morris, D. K., Ess, D. R., Hawkins, S. E., & Parsons, S. D. (1999). Development of a site-­specific
application system for liquid animal manures. Applied Engineering in Agriculture, 15(6),
633–638. https://doi.org/10.13031/2013.5829. https://elibrary.asabe.org/abstract.asp?aid=5829
Moshia, M. E., Khosla, R., Longchamps, L., Reich, R., Davis, J. G., & Westfall, D. G. (2014).
Precision manure management across site-specific management zones: Grain yield and eco-
nomic analysis. Agronomy Journal, 106(6), 2146–2156. https://doi.org/10.2134/agronj13.0400
Muha, I., Linke, B., & Wittum, G. (2015). A dynamic model for calculating methane emissions from
digestate based on co-digestion of animal manure and biogas crops in full scale German biogas
plants. Bioresource Technology, 178, 350–358. https://doi.org/10.1016/j.biortech.2014.08.060
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Niles, M.  T., Horner, C., Chintala, R., & Tricarico, J. (2019). A review of determinants for
dairy farmer decision making on manure management strategies in high-income countries.
Environmental Research Letters, 14(5). https://doi.org/10.1088/1748-­9326/ab1059
Pandey, A., Mai, V.  T., Vu, D.  Q., Bui, T.  P. L., Mai, T.  L. A., Jensen, L.  S., & de Neergaard,
A. (2014). Organic matter and water management strategies to reduce methane and nitrous
oxide emissions from rice paddies in Vietnam. Agriculture, Ecosystems and Environment, 196,
137–146. https://doi.org/10.1016/j.agee.2014.06.010
Paolini, V., Petracchini, F., Segreto, M., Tomassetti, L., Naja, N., & Cecinato, A. (2018).
Environmental impact of biogas: A short review of current knowledge. Journal of Environmental
Science and Health. Part A, Toxic/Hazardous Substances and Environmental Engineering,
53(10), 899–906. https://doi.org/10.1080/10934529.2018.1459076
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Pickering, N. K., Oddy, V. H., Basarab, J., Cammack, K., Hayes, B., Hegarty, R. S., Lassen, J.,
McEwan, J. C., Miller, S., Pinares-Patiño, C. S., & de Haas, Y. (2015). Animal board invited
review: Genetic possibilities to reduce enteric methane emissions from ruminants. Animal,
9(9), 1431–1440. https://doi.org/10.1017/S1751731115000968
3  Manure Management to Reduce Methane Emissions 45

Pierce, F. J., & Nowak, P. (1999). Aspects of precision agriculture. Advances in Agronomy, 67,
1–85. https://doi.org/10.1016/S0065-­2113(08)60513-­1
Pingak, G. M. F., Sutanto, H., Akhdiya, A., & Rusmana, I. (2014). Effectivity of methanotrophic
bacteria and Ochrobactrum anthropi as biofertilizer and emission reducer of CH4 and N2O
in inorganic paddy fields. Journal of Medical and Biological Engineering, 3(3). https://doi.
org/10.12720/jomb.3.3.217-­221
Prasanna, R., Kumar, V., Kumar, S., Yadav, A. K., Tripathi, U., Singh, A. K., Jain, M. C., Gupta,
P., Singh, P. K., & Sethunathan, N. (2002). Methane production in rice soil is inhibited by cya-
nobacteria. Microbiological Research, 157(1), 1–6. https://doi.org/10.1078/0944-­5013-­00124
Rabii, A., Aldin, S., Dahman, Y., & Elbeshbishy, E. (2019). A review on anaerobic co-digestion
with a focus on the microbial populations and the effect of multi-stage digester configuration.
Energies, 12(6). https://doi.org/10.3390/en12061106
Rad, A. K., Zarei, M., Astaikina, A., Streletskii, R., & Etesami, H. (2022a). Chapter 1 - effects of
microbial inoculants on growth, yield, and fruit quality under stress conditions. In M. Seymen,
E. S. Kurtar, C. Erdinc, & A. Kumar (Eds.), Sustainable Horticulture (pp. 1–38). Academic.
https://doi.org/10.1016/B978-­0-­323-­91861-­9.00014-­8
Rad, A. K., Astaykina, A., Streletskii, R., Afsharyzad, Y., Etesami, H., Zarei, M., & Balasundram,
S. K. (2022b). An overview of antibiotic resistance and abiotic stresses affecting antimicrobial
resistance in agricultural soils. International Journal of Environmental Research and Public
Health, 19(8). https://doi.org/10.3390/ijerph19084666
Rad, A., Zarei, M., Pourghasemi, H. R., & Tiefenbacher, J. P. (2022c). Chapter 27 – The COVID-19
crisis and its consequences for global warming and climate change. In H.  R. Pourghasemi
(Ed.), Computers in earth and environmental sciences (pp.  377–385). Elsevier. https://doi.
org/10.1016/B978-­0-­323-­89861-­4.00006-­3
Ramin, M., & Huhtanen, P. (2013). Development of equations for predicting methane emis-
sions from ruminants. Journal of Dairy Science, 96(4), 2476–2493. https://doi.org/10.3168/
jds.2012-­6095
Reay, D. S., Smith, P., Christensen, T. R., James, R. H., & Clark, H. (2018). Methane and global
environmental change. Annual Review of Environment and Resources, 43(1), 165–192. https://
doi.org/10.1146/annurev-­environ-­102017-­030154
Rehman, A., Ma, H., Irfan, M., & Ahmad, M. (2020). Does carbon dioxide, methane, nitrous
oxide, and GHG emissions influence the agriculture? Evidence from China. Environmental
Science and Pollution Research International, 27(23), 28768–28779. https://doi.org/10.1007/
s11356-­020-­08912-­z
Reinelt, T., & Liebetrau, J. (2020). Monitoring and mitigation of methane emissions from pressure
relief valves of a biogas plant. Chemical Engineering and Technology, 43(1), 7–18. https://doi.
org/10.1002/ceat.201900180
Ruile, S., Schmitz, S., Mönch-Tegeder, M., & Oechsner, H. (2015). Degradation efficiency of
agricultural biogas plants—A full-scale study. Bioresource Technology, 178, 341–349. https://
doi.org/10.1016/j.biortech.2014.10.053
Schellberg, J., Hill, M. J., Gerhards, R., Rothmund, M., & Braun, M. (2008). Precision agricul-
ture on grassland: Applications, perspectives and constraints. European Journal of Agronomy,
29(2–3), 59–71. https://doi.org/10.1016/j.eja.2008.05.005
Scheutz, C., & Fredenslund, A. M. (2019). Total methane emission rates and losses from 23 biogas
plants. Waste Management, 97, 38–46. https://doi.org/10.1016/j.wasman.2019.07.029
Schieffer, J., & Dillon, C. (2013). Precision agriculture and agro-environmental policy. In
J.  V. Stafford (Ed.), Precision agriculture 2013 (pp.  755–760). Wageningen Academic
Publishers. https://doi.org/10.3920/978-­90-­8686-­778-­3_93
Shamshiri, R. R., Balasundram, S. K., Rad, A. K., Sultan, M., & Hameed, I. A. (2022). An over-
view of soil moisture and salinity sensors for digital agriculture applications. In R. R. Shamshiri
& S.  Shafian (Eds.), Digital agriculture, methods and applications. IntechOpen. https://doi.
org/10.5772/intechopen.103898
46 A. K. Rad et al.

Stuart, D., Schewe, R.  L., & McDermott, M. (2014). Reducing nitrogen fertilizer application
as a climate change mitigation strategy: Understanding farmer decision-making and poten-
tial barriers to change in the US. Land Use Policy, 36, 210–218. https://doi.org/10.1016/j.
landusepol.2013.08.011
Subramanian, K. A., Mathad, V. C., Vijay, V. K., & Subbarao, P. M. V. (2013). Comparative evalu-
ation of emission and fuel economy of an automotive spark ignition vehicle fuelled with meth-
ane enriched biogas and CNG using chassis dynamometer. Applied Energy, 105, 17–29. https://
doi.org/10.1016/j.apenergy.2012.12.011
Tauseef, S. M., Premalatha, M., Abbasi, T., & Abbasi, S. A. (2013). Methane capture from live-
stock manure. Journal of Environmental Management, 117, 187–207. https://doi.org/10.1016/j.
jenvman.2012.12.022
VanderZaag, A.  C., Flesch, T.  K., Desjardins, R.  L., Baldé, H., & Wright, T. (2014).
Measuring methane emissions from two dairy farms: Seasonal and manure-management
effects. Agricultural and Forest Meteorology, 194, 259–267. https://doi.org/10.1016/j.
agrformet.2014.02.003
Wang, Z., & Calderon, M.  M. (2012). Environmental and economic analysis of application of
water hyacinth for eutrophic water treatment coupled with biogas production. Journal of
Environmental Management, 110, 246–253. https://doi.org/10.1016/j.jenvman.2012.06.031
Wang, Z. P., Lindau, C. W., Delaune, R. D., & Patrick, W. H. (1993). Methane emission and entrap-
ment in flooded rice soils as affected by soil properties. Biology and Fertility of Soils, 16(3),
163–168. https://doi.org/10.1007/BF00361401
World Bank. (2020). Climate watch. GHG emissions. World Resources Institute. Retrieved July
17, 2022. https://data.worldbank.org/indicator/EN.ATM.METH.KT.CE?end=2019&start=199
0&type=shaded&view=chart
Xiaohong, Z., Jia, H., & Junxin, C. A. O. (2011). Study on mitigation strategies of methane emis-
sion from rice paddies in the implementation of ecological agriculture. Energy Procedia, 5,
2474–2480. https://doi.org/10.1016/j.egypro.2011.03.425
Yacob, S., Hassan, M. A., Shirai, Y., Wakisaka, M., & Subash, S. (2005). Baseline study of meth-
ane emission from open digesting tanks of palm oil mill effluent treatment. Chemosphere,
59(11), 1575–1581. https://doi.org/10.1016/j.chemosphere.2004.11.040
Yusuf, R.  O., Noor, Z.  Z., Abba, A.  H., Hassan, M.  A. A., & Din, M.  F. M. (2012). Methane
emission by sectors: A comprehensive review of emission sources and mitigation methods.
Renewable and Sustainable Energy Reviews, 16(7), 5059–5070. https://doi.org/10.1016/j.
rser.2012.04.008
Zamorano, M., Ignacio Pérez Pérez, J., Aguilar Pavés, I., & Ramos Ridao, Á. (2007). Study of the
energy potential of the biogas produced by an urban waste landfill in Southern Spain. Renewable
and Sustainable Energy Reviews, 11(5), 909–922. https://doi.org/10.1016/j.rser.2005.05.007
Zarei, M., & Kaviani Rad, A. K. (2020). Covid-19, challenges and recommendations in agricul-
ture. Journal of Botanical Research, 2(1), 12–15. https://doi.org/10.30564/jrb.v2i1.1841
Zhang, A., Cui, L., Pan, G., Li, L., Hussain, Q., Zhang, X., Zheng, J., & Crowley, D. (2010). Effect
of biochar amendment on yield and methane and nitrous oxide emissions from a rice paddy
from Tai Lake plain, China. Agriculture, Ecosystems and Environment, 139(4), 469–475.
https://doi.org/10.1016/j.agee.2010.09.003
Chapter 4
Crop Residue Incorporation to Enhance
Soil Health in the Rice–Wheat System

Hamna Bashir, Waqas Mohy-Ud-Din, Zahoor Mujdded Choudary,


Muhammad Mahroz Hussain, and Muhammad Ashir Hameed

Abstract  There has been a fourfold rise in the world population in the past century.
To feed the ever-increasing numbers of people, increased agricultural and industrial
processes have put further burden on food production. The utilization of crop waste
in fields might be regarded as crucial in developing countries. Agricultural soil
health is altered for increasing physical, chemical, and biological process owing to
the lack of alternative organic amendments. Agricultural residual management tech-
niques in developing countries, i.e., surface retention, integration, and removal, are
discussed in this chapter with their benefits and hazards to the agroecosystems
based on cereal crops. The health of agricultural soils has deteriorated as a result of
increased food production over time. Nutrient cycling and soil quality are influ-
enced due to the various biological, chemical, and physical processes that occur on
organic matter that is returned to the soil in the form of crop residues. This chapter
will discuss important biological properties like soil microbial biomass and soil
biodiversity; physical properties like soil moisture content, soil temperature, soil
compaction, and erosion; and chemical properties like soil cation exchange capac-
ity, soil pH, and soil organic carbon. The competitive use between residue retention
and yield in mixed crop/livestock systems in developing nations can be a problem.
On the other hand, strategies such as intensification and partial retention, as well as
nutrient cycling from manures and alternatives to the current functions of livestock
such as a mechanized system or insurance, could reduce the pressure on residues in
favor of long-term soil quality and health.

Keywords  Soil health · Physicochemical attributes · Nutrient cycling · Crop


residues

H. Bashir · W. Mohy-Ud-Din · Z. M. Choudary · M. M. Hussain (*) · M. A. Hameed


Institute of Soil and Environmental Sciences, University of Agriculture Faisalabad,
Faisalabad, Pakistan

© The Author(s), under exclusive license to Springer Nature 47


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_4
48 H. Bashir et al.

1 Introduction

There has been a fourfold rise in the world population. To feed the ever-increasing
number of people, increased agricultural and industrial processes have put further
tension on the production of food. In the past years, increased food production has
caused serious damage on agricultural soil health and quality (Dias et  al., 2015;
Kareem et al., 2022). In different regions of Africa, Asia, and Latin America, crop
yields have decreased significantly due to the increase of soil degradation (Din
et al., 2022; Alewell et al., 2020). According to the Food and Agriculture Organization
(FAO), soil health is referred to the capability of the soil to function as a living sys-
tem and fitness for use refers to the quality of soil (Bünemann et al., 2018). A good
soil quality has less deterioration and is thus very productive in agriculture produc-
tion (Eswaran et al., 2019). Soil health has an impact on soil quality, which is crucial
for long-term agricultural output. It is important to remember that soil is a living
system, which means it contains organisms that perform a variety of functions such
as the recycling of nutrients; controlling pests, weeds, and diseases; maintaining
symbiosis with roots; improving soil gaseous exchange; and enhancing soil aggre-
gate formation. Soils with high organic matter content are home to thriving soil
organisms, which serve as a reservoir for nutrients and moisture (Table 4.1). Organic
amendments must be added to the soil regularly to enhance or maintain its organic
matter content, which in turn improves soil health (Hussain et al., 2021c; Urra et al.,
2019; Parray et al., 2022; Bandh et al., 2022; Bandh, 2022a, b).
Crop residue (the remains of harvesting) is the most readily available biomass.
Zhou et al. (2016) claim that crop residues are the biggest source of soil organic
matter (SOM). Rice (Oryza sativa L.), sorghum (Sorghum bicolor L.), maize (Zea
mays L.), and wheat (Triticum aestivum L.) are among the primary cereal crops that
produce considerable volumes of agricultural waste (Aula et  al., 2019). In 2010
approximately 217 Mha for wheat, 161 Mha for maize, 154 Mha for rice, and 41
Mha for sorghum were harvested across the world (Zuo et al., 2018). Twenty-five
percent of maize and wheat crops are sources of calories for the people living in
developing countries, which together account for 40% of all food eaten globally
(Wijesinha-Bettoni & Mouillé, 2019). Retaining agricultural waste or incorporating
it into the soil improves soil quality in several ways (Farooqi et al., 2021). Although
small-scale farmers in underdeveloped countries confront a trade-off in controlling
agricultural waste, in certain cases, crop residues can be used as biofuel and animal
feed, or they can be grazed in crop fields by livestock. There are several ways farm-
ers use to prepare their fields for sowing one of them is burning off crop waste. The

Table 4.1  Essential plant nutrients in soil


Nutrient type Nutrients
Primary nutrients Nitrogen, phosphorus, and potassium
Secondary nutrients Calcium, magnesium, and sulfur
Macronutrients Boron, chlorine, copper, iron, manganese, molybdenum, nickel, and zinc
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 49

long-term environmental and economic advantages of conserving agricultural


wastes need a shift in conventional crop residue management.
The rice–wheat cropping system (RWS) is one of the most frequently employed
cropping systems in India, with around 90% of the land used in the Indo-Gangetic
Plains (IGPs) (Singh & Sidhu, 2014). Combine harvesters have made it possible to
mechanically harvest more than 75% of the rice land in the northwestern IGPs.
Wheat straw is often used by farmers to feed their livestock. Due to its high silica
concentration, rice straw is regarded as a poor feed for livestock, making manage-
ment of huge biomass considerably difficult. The seed drill used to sow wheat is
hampered by a swath of loose rice residues left behind by the combine harvester.
Farmers burn agricultural leftovers to prevent these issues (90–140 Mt annually).
Rice straw can be disposed of by burning, according to the perspective of farmers.
Not only does it save money, but it also serves as an excellent form of pest manage-
ment (Minas et al., 2020). India’s 0.05% greenhouse gas emissions are attributable
to the burning of rice straw, according to a study by Bisen and Rahangdale (2017),
whose results showed significant loss of biomass and have a negative impact on soil
fauna and flora and also in different soil characteristics (Malla et al., 2022; Bandh
et al., 2021, 2023; Mushtaq et al., 2020).

2 Crop Residue Impact on Soil Physiochemical Health

Crop residue is returned to the soil in the form of organic matter that undergoes dif-
ferent biological, chemical, and physical processes that all work together to influ-
ence soil quality and the cycling of nutrients. Soil organic carbon, soil pH, and
cation exchange capacity are all affected by residue management (Hussain et al.,
2022); physical qualities, e.g. soil compaction and moisture content; and biological
properties, e.g., soil biodiversity and microbial biomass (Farooqi et  al., 2022).
Because increased crop yields leave more agricultural residue after harvest, crop
yield findings can be provided in this study because of their contribution to the post-
harvest residue of crops (Fig. 4.1).

2.1 Structure of Soil

The resilience of a system based on crop production, soil erosion, and degradation
is strongly influenced by the structure of the soil. Rainfall, tillage, mechanization,
and residue management all have an impact on the physical stability of soil organic
matter and soil structure (Sithole et al., 2016; Turmel et al., 2015). Soil structure can
be improved by crop residues in many ways: by boosting soil aggregation via the
organic matter added to topsoil and by avoiding the soil compaction due to rain-
drops (Almendro-Candel et al., 2018). The structural stability of soil can be calcu-
lated by looking at how long soil aggregates can hold together under pressure. The
50 H. Bashir et al.

Fig. 4.1  Simplified model of plant residue inputs transformed by soil microorganisms. (Reprinted
with permission from Turmel et al. (2015))

stability of soil aggregates affects soil porosity, water, gas, nutrient flow in the soil
system, and root growth (Almendro-Candel et al., 2018). To keep soil organic mat-
ter from decomposing, soil aggregates can create organo-mineral complexes that
are incomprehensible to microorganisms (Havlicek & Mitchell, 2014).
There are several reasons why soil aggregates are important, but the loss of soil
organic matter (SOM) is a major concern, especially in agricultural soils in specific
surroundings where SOM is widespread. In mountainous terrains, intense rainfall,
and extensively eroded soils, agricultural residues can play an essential role in sta-
bilizing topsoil aggregates. In a 6-year field experiment on sandy loam soil,
Bhattacharyya et al. (2012) compared various tillage strategies with the crop remains
integrated or left on the topsoil in a lentil-finger millet cycle (Lens esculentus
L.–Eleusine coracana L.): no-tillage (NT) and one no-tillage on seasonal basis;
conventional tillage (NT–CT) treatments, where residues of the crops were left on
the soil surface and had higher water-stable macroaggregates in the surface soil
layer (0–5  cm) than year-round conventional tillage (CT–CT); and one-seasonal
conventional tiling (CT–NT) treatments, where residues were incorporated. In the
NT–CT and NT treatments, higher soil organic carbon (SOC) concentrations were
due to the soil compaction. Due to larger crop residual biomass on the surface, the
absorption and breakdown of biomass in the absence of tillage resulted in delayed
formation of SOC in the surface soil layer, resulting in a more stable aggregate and
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 51

SOC buildup (Prasad et al., 2016). As long as residue remains on the soil surface,
NT and seasonal NT–CT techniques are effective for aggregate stability and pre-
serving soil organic carbon in soils that are prone to soil erosion (Bhattacharyya
et al., 2012). Better soil aggregate stability was observed when plants retained their
residues on the soil surface (Turmel et al., 2015). This might be due to an increase
in microbial activity and adhesive agent production caused by organic N from the
residues.
No-till systems are better at keeping residues on the surface than absorbing them,
which can raise soil temperature and lead to increased mineralization of residues
(Vanhie et al., 2015). Tillage has the potential to be a more critical element than resi-
due addition in driving aggregate formation. Central China is famous for rice culti-
vation having clay-loam soil, and also famous for no-tillage practices, it was found
that in surface soils under continuous NT, with or without, the crop residues of
rape–rice rotation had significantly higher proportions of water-stable aggregates
and particulate organic C than in single or continuous CT without residue. When
organic materials remained undisturbed, even in the NT, it helped to promote mac-
roaggregate formation and residue retention (Li et al., 2012). Soil aggregate stabil-
ity has been increased due to the residue retention on topsoil; however, findings in
subsoils under tillage regimes with and without residue retention have proven con-
tradictory (Hameed et al., 2021b). No-tillage with or without residue reduced mac-
roaggregate proportions from the 5 to 30  cm layer, but in conventional ways of
tilling, the treatments without crop residue increased macroaggregate fractions from
5 to 15 cm to the 5 to 30 cm layer under NT (Bhattacharyya et al., 2012). Chen et al.
(2017) reported that there is a rise in the percentage of macroaggregates in the soil
plow layer, while others have found no major variations in the subsurface soil layer
when using reduced tillage processes (Bhattacharyya et al., 2012). Because external
inputs (seed, fertilizer, and pesticides) are applied to this horizon, macroaggregate
proportions can vary from soil to soil, but it is apparent that for the better production
of the crop, the soil surface layer is very important (Hussain et al., 2021a).
Surface organic matter has an important function in safeguarding topsoil aggre-
gates, for the preservation of soil erosion, permitting water penetration, and nutrient
retention since it is the soil layer that is more prone to rain and connects the atmo-
sphere to the soil (Menta, 2012). Soil aggregates are protected from raindrop impact
by the preservation of surface residue. Dissipating rainfall’s influence and keeping
soil aggregates intact are two of the most essential functions of residue cover
(Turmel et al., 2015). Irrigation water is becoming limited in the Yaqui Valley in
Sonora, Mexico, where surface debris is fed to livestock or tilled into the soil. A
study by Verhulst et al. (2011a) found that the soil aggregate stability of the 0–5 cm
layer was less in traditionally raised beds and more in permanent raised beds because
of the protection provided by the surface residue from raindrop impact. Furthermore,
the soil is protected from compaction by leaving residues on the surface. Water
infiltration in soil, air movement, and soil porosity are all negatively affected by the
compaction of the soil. When soil is compacted, nitrogen availability and crop
development can be negatively affected, as well as loss of nutrients via surface run-
off (Kaur et  al., 2020). Residue management practices, equipment usage,
52 H. Bashir et al.

experimental plan, and soil texture all influence the impact of agricultural residue
on soil compaction (Celik et al., 2017).
The compaction of surface soil can rise dramatically over many years of contin-
ual NT if no residue is left behind. Runoff and erosion can occur as a result of com-
paction because of the seal-over and crusting. Soils with zero tillage and burned
crop waste showed the greatest penetration resistance in northwest Mexico (Verhulst
et al., 2011b). It was shown that infiltration rates under NT were greater when resi-
due was retained on the surface without residue retention as compared to NT. Turmel
et al. (2015) demonstrated that surface residue can help prevent soil compaction’s
harmful impacts. Under mechanical no-till, in clayey soils, compaction of soil can
be a prominent problem. There are subsoils containing gravel and clay with low
organic matter content and a potential to crust on the top of alfisols in many loca-
tions of Africa where crops like sorghum and maize are produced, making alfisols
susceptible to compaction. No-till and residue retention have been proven to be
effective methods of minimizing soil compaction in Africa’s humid and subhumid
areas (Somasundaram et  al., 2020). Penetration resistance and bulk density were
reduced in no-till with mulch in tropical alfisol in Ibadan, Nigeria, for example, as
compared to no-till without mulch (Fasinmirin & Reichert, 2011). Soil residue
retention is critical in no-tillage systems for reducing soil compaction.

2.2 Impact of Crop Residues on Organic Carbon

Crop residues increase soil aggregate stability and water retention and also act as a
nutrient reservoir. Researchers consider soil organic matter as a good indicator for
soil quality and agricultural stability (Wander et al., 2019). Carbon (C) is emitted as
CO2 as a result of soil heterotrophic and autotrophic respiration. Although the use of
land changes with the passage of time, the disposal of crop residues and digestion
process in animal produce more CO2 as compared to human activities (Ali et al.,
2019). Different techniques in agriculture disrupt the soil organic carbon pool,
which is a significant source of greenhouse gas; carbon losses in soil decrease pres-
sure on long-term crop yield, food security, and soil quality. Agricultural practices
can either increase inputs of organic matter or can also delay the rate of breakdown
of cellulose to maintain SOC levels (Mitchell et  al., 2018). Soil organic matter
(SOM) can be categorized in two ways: humus pool and labile pool, both influenc-
ing nutrient availability and carbon storage. Humus is resistant to degradation,
whereas labile pool is easily degradable by soil microorganisms, resulting in more
carbon storage in the soil. The physical and chemical stabilization of this pool
ensures its long-term stability (Dheri & Nazir, 2021). Decomposition of crop resi-
due is an important step in the process of humus formation, which results in carbon
storage.
Retention of crop residues is critical for raising and/or sustaining levels of SOC,
although soil type, climate, and management techniques can influence its impact
(Turmel et  al., 2015). The conservation of mulch ripping residue, as opposed to
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 53

clean ripping with residue eliminated, resulted in increased soil organic carbon in
sandy soils in Zimbabwe (Nyakudya & Stroosnijder, 2015). No major changes in
soil organic carbon were observed in sandy soil and red clay soil. In addition to
influencing breakdown rates, climatic conditions can have an impact on the quantity
of SOC that can accumulate with residue remaining on the surface rather than being
incorporated. Turmel et al. (2015) found that rain-fed maize cultivation in Apatzingán
and Casas Blancas, Mexico, resulted in organic C sequestration for 6 years. Soil
organic carbon content in Casas Blancas and Apatzingán was found to be higher
with four treatments including minimum tillage, NT with 33% crop residues, 66%
crop residues, and 100% crop residues by comparing conservational tillage treat-
ment having disk plowing and disking with 0% crop residue NT. Surface decompo-
sition of crop residue under no-tillage is anticipated to be slower than in normal
tillage when they are mixed with soil and come in contact with soil microorganisms
(Wang et al., 2020). Soil organic carbon and total nitrogen (N) were greatest when
residue was left on the surface rather than incorporated under minimum tillage in
Varanasi, India (Turmel et al., 2015).
The retention of crop residues in the soil profile can also be influenced by man-
agement practices such as surface retention of crop residues or tillage incorporation.
Conventional tillage is often blamed for increased decomposition rates, which result
in C losses (West & Six, 2007). When soil structure is disturbed and organic matter
is redistributed, tillage releases carbon due to surface microbial activity (Sapkota,
2012). This approach, cultivating agricultural soils, has resulted in a 30–50%
decrease in pre-cultivation SOC (Bruun et al., 2015). Hijbeek et al. (2017) stated in
their meta-analysis of soil carbon case studies quite conflicting results. There was
no significant change in soil C stock between zero tillage and conventional tillage in
7 out of the 78 research studies. Understanding decomposition and SOM stabiliza-
tion is essential for determining the fate of carbon and nutrients via crop waste and
other soil amendments, delivered to the soil when selecting appropriate manage-
ment techniques (Navarro-Pedreño et al., 2021). No-tillage with residue retention
has shown a greater SOC content. However, this concentrates C on the soil surface
(Luo et al., 2020). According to Zhang et al. (2018), soil tillage affects SOC distri-
bution by integrating residues into the soil and therefore raising SOC in deeper
layers. All SOC can be kept in a single layer of soil when the whole soil profile is
examined (Powlson et al., 2011). The SOC content was highest in the 0–5 cm soil
layer without surface residue for conventional tillage, with incorporated chopped
surface residue rotary tillage, and with standing residue for no-tillage in the Northern
China Plain silt loam soils (Mert et al., 2018).
However, up to 30 cm SOC content decreased with depth. The 5–10 cm deep soil
layer had the greatest SOC concentration under conventional tillage with residue
treatment. Even while moldboard plowing is often demonstrated to reduce C stocks,
conventional tillage with residue treatment increased significantly in this research,
illustrating once again the varying influence on SOC induced by crop residue man-
agement strategies. Therefore, shallow sampling does not favor no-till methods
such as residue retention and provides more accuracy in the evaluation of the effec-
tive management strategies of residues on soil organic carbon, sampling the whole
54 H. Bashir et al.

plow depth (Yang et al., 2013). In addition, when comparing residue surface reten-
tion and incorporation outcomes, various parameters (e.g. soil, time, experiment,
environment) should be considered (Powlson et al., 2011). There is no doubt that
no-till in conservation agriculture with surface residue has its advantages, but it is
unclear whether they are better at carbon sequestration than the incorporation of
residues or tillage, and soil sampling methods can give better results compared with
the direct effect of soil organic carbon storage (Valkama et al., 2020). Experimental
studies in temperate areas account for the bulk of the data, and knowledge on this
topic is especially scarce in tropic and subtropic regions (Deryng et al., 2014).

2.3 Percentage of Different Bioavailable


Micro- and Macronutrients

Nitrogen availability can be affected by the addition of agricultural wastes. Nitrogen


(N) mineralization may occur when the crop residue of legume having a low C/N
ratio is added, but during the breakdown with a high C/N ratio, legume residues
partially immobilize N.  Increase of soil moisture content and improper fertilizer
incorporation and larger denitrification losses may occur from chemical N fertiliz-
ers when the crop residues remain on the soil surface (Walsh & Belmont, 2015).
There is evidence that retaining residue in the topsoil increases the content of
phosphorus (P). There may be movement of phosphorus from lower layers in the
soil (Iqbal et al., 2011). The heavily weathered soils and the addition of residues can
have an indirect effect on phosphorus availability. Humic compounds and aliphatic
acids with low molecular weight generated can block Al oxide adsorption sites dur-
ing the degradation of agricultural wastes and minimize P adsorption (Johan et al.,
2021; Qadir et al., 2021). In general, legumes are more effective because of their
higher decomposition rates; this impact depends on the quality of the residue (Abera
et al., 2012).

2.4 Crop Residue Effect on Soil pH

Soil pH can be affected by crop residue retention because of the relationship between
crop residue, chemical composition, and soil characteristics (Turmel et al., 2015).
The concentration of nitrogen and alkaline ash in legume crop residues is more
likely to affect pH than in wheat crop residues (Butterly et al., 2013). The impact of
residue incorporation on soil’s initial pH can be affected due to different soil char-
acteristics including temperature, moisture, and soil organic carbon availability that
can pace the residual decomposition (Lehman et al., 2015). The anionic concentra-
tion in organic form and the N content in crop residues are connected with changes
in pH as a result of crop residue addition (Turmel et  al., 2015). Organic anionic
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 55

decomposition consumes H+ ions as a result of soil pH changes; H+ ion is associated


with organic ions if the initial soil pH is lower than the pKa value of a weak acid
group of organic carbon and vice versa (Adusei-Gyamfi et al., 2019).
The ammonification and nitrification process both release and consume H+
throughout the nitrogen mineralization process, resulting in pH shifts in the soil.
The overall effect is acidifying due to a state of disequilibrium in the nitrogen cycle.
Final soil pH is determined by the reaction balance (Heil et al., 2016). Incorporation
vs. surface retention of residue can have an impact on pH in the topsoil layer,
although this remains a matter of debate. With the incorporation of crop residues of
maize and wheat in topsoil in central Mexico for 5 years, the surface pH was 7.0 and
without residues 6.6 and 6.7 when residues are incorporated with tillage (Dendooven
et  al., 2012). However, results were not significant for subsoils with depths of
5–20 cm. Sombrero and De Benito (2010) observed the comparison between chisel
plow minimal tillage and conservation tillage and no-tillage with incorporated resi-
due and para low zone tillage in central Spain. The first two operations lowered the
pH than the second two operations. When maize residues were integrated into the
soil, Turmel et al. (2015) found a decrease in pH in both treatments. The pH in the
0–10 cm layer was initially 6.7 but declined in the first 3 years of the experimental
studies before stabilizing at 5.6 when residues were maintained on the surface and
5.4 when they were integrated into the soil.

3 The Activity of Soil Microbes

Soil microbes decompose the organic material that affects soil structure and water
and nutrient availability in agroecosystems. The living part of soil organic matter is
termed soil microbial biomass (SMB) (Naveen Kumar & Babalad, 2018). Soil qual-
ity can be improved and considered a sensitive and valuable indicator due to its role
as a source and reservoir of biologically accessible nutrients and as a catalyst for
soil aggregation and structure formation (Huera-Lucero et al., 2020). Temperature
and moisture in the environment, as well as soil management methods such as resi-
due inputs, can alter SMB (Tiwari et  al., 2019). Residue retention stimulates the
microbial activity and SMB. Zhang et al. (2016) found that when straw was retained
rather than removed, microbial biomass C levels were significantly higher because
of increased soil porosity, soil moisture, and decreased temperature of soil due to
the residue cover in North China. Soil microbial biomass levels were shown to be
greater in experiments conducted in central Mexico when residue was incorporated
as opposed to when residue was removed (Verhulst et al., 2011b).
Adding of crop residues to the soil warms it and aerates it more quickly, which
benefits microorganisms and speeds up decomposition. This results in a larger loss
of soil organic carbon (SOC) over time (Thangarajan et al., 2013; Victoria et al.,
2012). Experimental studies showed that changes in agricultural residual supply
altered the biomass of microorganisms, which is consistent with this finding (Feng
et  al., 2022; Li et  al., 2018). Soil organic carbon has been increased due to the
56 H. Bashir et al.

incorporation of crop residues and no-till with high temperatures and rainfall, com-
pared to the incorporation of SOC (Chowdhury et al., 2015). Less interaction occurs
between soil microorganisms and crops residing on the surface than in humid tem-
perate or humid tropical climates where breakdown rates are high, suggesting the
importance of residue preservation on the soil surface rather than incorporating
them into the soil. Arbuscular mycorrhizal fungi (AMF) are an example of a fungus
that can enhance plant nutrient availability. Symbiotic connections exist between
AMF and plant roots, with the AMF obtaining nutrients from the plant and the plant
roots receiving phosphate through hyphae (Basu et  al., 2018). Additionally, soil
particles get aid from the glycoprotein and AMF hyphae for the bonding that results
in improving the stability of the aggregates (Bronick & Lal, 2005; Li et al., 2022).
Development of mycorrhiza hyphae is adversely affected by tillage practices, and
the organic matter addition to soil has been shown to increase AMF growth and
spore production (Wei et al., 2019).

4 The Activity of Earthworms

Macrofauna, such as earthworms, play a significant role in the soil ecosystem as


well. Soil ecosystems can benefit from the long-term benefits of earthworms, which
are referred to as ecosystem engineers (Bender et al., 2016). Their biomass directly
affects C and N cycles by taking in and storing nutrients, as well as releasing con-
siderable quantities of N via excretion (Mischler et  al., 2016). Incorporation of
organic matter into the soil affects aggregate stability and C and N cycles through
their digestive tracts as well as their middens, casts, and burrows. These stimulate
microbial processes like breakdown and mineralization by bringing microorgan-
isms into direct contact with organic materials (Paterson & Sim, 2013). Retention of
crop residues and little soil disturbance have been shown to encourage earthworm
behavior. Surface crop residues that are left behind after harvest provide food sup-
ply and lower the soil temperature, all of which can contribute to an increase in the
biomass of earthworms and their populations (Abail & Whalen, 2018). On the other
hand, tillage has been demonstrated to have a detrimental effect on several species
of earthworms because it destroys their burrows, giving them physical injury and
exposing them to the surface for the predator (Briones & Schmidt, 2017). Soil inver-
tebrates, particularly earthworms, show an increase in their population and variety
in Tunisia (semiarid region) owing to the improvement of soil characteristics and
the absence of disturbance. Tillage can assist endogeic (horizontal-burrowing)
earthworms, providing a food supply, if the residue is integrated into the soil, as
residue retention can have a varied influence on earthworms (Turmel et al., 2015).
The incorporation of residue rather than spreading it on the soil surface might limit
their population even with shallow tillage in areas with high populations of anecic
(vertical burrowing) earthworms because they remain on the surface and used as a
food source (Bertrand et al., 2015).
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 57

Bertrand et  al. (2015) also discovered an earthworm population connection


between crop and tillage. Under crop field earthworm numbers were low in conven-
tional and reduced tillage treatments, but earthworm numbers were high in all treat-
ments under the grass–clover cycle. Crop residue impact on soil fauna and
earthworms might depend upon the plowing depth, incorporation of residues, type
of crop residue, quality of residue, frequency of plowing, and crop residue type
(Mirzaei et al., 2021; Younas et al., 2021).

5 Methods for Crop Residue Management

India’s rice–wheat farming system creates an enormous quantity of agricultural


waste, as one would expect. In northwest India, combine harvesters are used to
harvest rice–wheat crops, which leave residues. Cattles are the primary consumers
of cereal crop residues. Domestic stoves and rice parboiling boilers use rice straw
and husk as a fuel. It is more difficult to manage rice straw than wheat straw because
of the shorter turnaround time of wheat–rice crops, the lack of appropriate recycling
technology, and rice straw’s higher silica content. Incorporation of biochar, baling,
mulching, and removing the straw are the management operations available to farm-
ers for profitably managing crop residues. Straw management practices vary
depending on the situation.

5.1 Animal Feed from Crop Residues

In India, the traditional use of crop waste is to feed it to animals and supplement
it with various additives. It is not possible to use crop wastes as a single feed for
cattle because of their poor digestion. Rice straw is considered poor feed for
animals due to its high silica content. It has low lignin content than other crop
straws. It is possible to increase the nutritional content of rice straw by several
approaches. Crop residues’ lignocellulose linkages have been weakened and bro-
ken down by physical, chemical, and biological treatments (Ginni et al., 2021).
As food for livestock, wheat straw is mostly shred into small pieces with a cut-
ting machine. However, this process demands extra time and resources. Because
rice straw stems have a lower silica content than leaves, cutting the rice crop as
close to the ground as possible improves the digestibility of the straw for live-
stock. Urea and molasses can be added to the residues, and green fodders includ-
ing leguminous and nonleguminous crops can be used to meet the nutritional
needs of animal feed.
58 H. Bashir et al.

5.2 Biofuels from Crop Residues

Reduced reliance on fossil fuels via the use of biofuels is unquestionably an essen-
tial option. This process is critical because ethanol can be used as a fuel extender
and octane-enhancing agent in gasoline, or it can be used as a clean fuel in internal
combustion engines. Different feedstock (corn cob, rice straw, wheat straw, bagasse,
and sawdust) have theoretical ethanol production estimates ranging from 382 to 471
liters per ton of dry matter. Ethanol production from agricultural waste is becoming
more and more sophisticated in India. The conversion of crop residues to alcohol
has a few limiting steps that need to be improved (Hameed et al., 2021a; Hussain
et al., 2021b).

5.3 Biochar from Crop Residues

As a realistic technique for sustaining soil health, biochar from crop residues has
gained a lot of attention in the few past years. Slow pyrolysis (heating without oxy-
gen) of biomass produces fine-grained charcoal with a high carbon content known
as biochar. It has the capacity to store carbon in soil over the long term. Biochar
made up to plant biomass is resistant to decomposition by soil microorganisms
when it is applied to the soil (Heikkinen et al., 2021; Hussain et al., 2022). Biochar
improves water quality via its ability to absorb impurities and minimize greenhouse
gases in agricultural areas. It is important to keep in mind that the characteristics of
biochar might change depending on the biomass used and the pyrolysis condi-
tions used.

5.4 Incorporation of Crop Residues

With the incorporation of crop residues into the soil instead of removing or
burning agricultural waste, straw integration enhances soil organic matter, as
well as the concentrations of nutrients N, P, and K in soil. Ball et al. (1990) and
Christian and Bacon (1991) found that plowing is the most efficient method of
incorporating residues into the soil (Farid Eltom et al., 2015). Remaining crop
leftovers might be absorbed into the soil partly or entirely depending on the
manner of cultivation. It is more difficult to incorporate rice residues into the
soil before wheat planting than to incorporate wheat straw into the soil before
rice planting. Adding crop residues to the field is a good way of nutrient recy-
cling, but it also immobilizes the soil nutrients temporarily (such as nitrogen),
which causes the additional burden of nitrogenous fertilizer use to adjust the
high C/N ratio (Turmel et al., 2015). The short-term immobilization of available
soil and fertilizer nitrogen by decomposer microbial activity is the source of this
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 59

N deficiency. The length of time depends on the decomposition rate of crop


waste, as well as the crop residue quality and the soil environment (Grzyb et al.,
2020). Similarly, Upadhyay et al. (2012) stated that in 10, 20, and 40 days of
wheat planting, no negative impact was observed on wheat and subsequent rice
grain yields from in situ incorporations of rice straw in soil. Addition of rice
straw to the wheat crop did not even have any influence on the next year’s rice
harvest.

5.5 Surface Mulching from Crop Residues

Evaporation of water can be minimized by retaining soil residues on the


earth’s surface and is a better option for soil conservation. Soil microbial
populations are increased, which in turn improves soil health by increasing
organic carbon. Soil health directly affects wheat production (Turmel et  al.,
2015). Zero-till wheat has been implemented in the rice–wheat system in the
northwest IGP.  For this, an advanced seed drill has been developed. Sidhu
et al. (2007) reported that in conservation agriculture any kind of residue can
be used with the Happy Seeder as long as it is spread out evenly before drill-
ing. Wheat grain production was up to 31% higher with rice straw mulch, crop
water usage was reduced to 31%, and water use efficiency was up to 25%
higher than it would have been without it. The density of root length in lower
levels of >0.15  m was 40% higher when mulch was used compared to no-
mulch, which can be attributable to the increased soil moisture retention in
lower layers (Singh & Sidhu, 2014).

6 Conclusions

Crop residue significantly increased the quality of soil in Asia, Latin America,
and Africa. Surface residue retention typically improves soil health qualities
physically, chemically, and biologically, but it has been shown to have detri-
mental impacts on crop production in specific conditions. Some studies have
shown that retaining residue, whether it is absorbed into the soil or left on the
top, is more beneficial than completely removing it or incorporating it into
intensive tillage operations. When comparing different residue management
approaches, it is crucial to consider abiotic parameters like the texture of the
soil, regional climate, research length, sample methodologies, and agricultural
activities like weed control as well as surface and residue retention vs. assimi-
lation. There are several advantages to incorporating the crop residue rather
than mixing it into the soil, including greater topsoil protection against soil
loss, erosion, and surface compaction. Soil erosion and rapid decomposition
rates in humid tropical environments necessitate promoting residue retention
60 H. Bashir et al.

on the surface in these situations. Soil is affected by crop residue and is largely
dependent on the quality of residues and parameters of the soil. While the
influence on SOC and CEC can only be seen on the soil surface, with the incor-
poration of residues into the soil, the effect can be seen at a larger depth.
Sample methods that compare residue management strategies should take this
into account. This results in enhanced microbial activity and faster breakdown
rates when residue is left in the ground. In addition to stimulating the earth-
worm population, residue retention has a corresponding impact on earthworm
ecology.
Crop residue retention has been shown to have adverse impacts in specific situ-
ations; however, these cases are rare. It is the most prevalent negative impact of N
mobilization that might lower the supply of N to the crop under N limiting condi-
tions. Strip tillage should also be used in colder areas since surface residues
reduce the soil’s temperature, which can severely affect the crop’s output. In areas
having significant rainfall, excess soil moisture retention can cause waterlogging.
On the other hand, leftovers in semiarid locations with little rainfall can act as rain
catchers and so accelerate the rate of evaporation. There is an exchange between
the use of crop residue incorporation and animal waste for the purpose of soil
quality. Crop leftover can be replaced with noncommercial plants used as mulch,
or as much as 30–60% or 50–75% of the crop residue can be left on the soil sur-
face, depending on the region’s soils and temperature. Because improved soil
quality and moisture retention increase yields in areas where water is available to
crops in limited quantities, retaining crop residue can have a positive feedback
effect, increasing crop residue that can be used as animal feed. It is possible to
further benefit from soil health by retaining leftovers on the surface, but farmers
in open range areas prefer to integrate their residues to keep them from being
grazed by their neighbors’ livestock.

Acknowledgements Authors want to thank the Higher Education Commission, Pakistan, for


awarding fellowship under the International Research Support Initiative Program (IRSIP)
Muhammad Mahroz Hussain no. 1-8/HEC/HRD/2020/10831, and also to the Environmental
Biogeochemistry Laboratory, University of Agriculture Faisalabad, Pakistan.

References

Abail, Z., & Whalen, J. K. (2018). Corn residue inputs influence earthworm population dynamics in
a no-till corn-soybean rotation. Applied Soil Ecology, 127, 120–128. https://doi.org/10.1016/j.
apsoil.2018.03.013
Abera, G., Wolde-Meskel, E., & Bakken, L.  R. (2012). Carbon and nitrogen mineralization
dynamics in different soils of the tropics amended with legume residues and contrasting
soil moisture contents. Biology and Fertility of Soils, 48(1), 51–66. https://doi.org/10.1007/
s00374-­011-­0607-­8
Adusei-Gyamfi, J., Ouddane, B., Rietveld, L., Cornard, J. P., & Criquet, J. (2019). Natural organic
matter-cations complexation and its impact on water treatment: A critical review. Water
Research, 160, 130–147. https://doi.org/10.1016/j.watres.2019.05.064
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 61

Alewell, C., Ringeval, B., Ballabio, C., Robinson, D. A., Panagos, P., & Borrelli, P. (2020). Global
phosphorus shortage will be aggravated by soil erosion. Nature Communications, 11, 1.
Ali, M., Saleem, M., Khan, Z., & Watson, I. A. (2019). The use of crop residues for biofuel produc-
tion. In Biomass, biopolymer-based materials, and bioenergy (pp. 369–395). Elsevier.
Almendro-Candel, M.  B., Lucas, I.  G., Navarro-Pedreño, J., & Zorpas, A.  A. (2018). Physical
properties of soils affected by the use of agricultural waste. Agricultural Waste and Residues,
2, 9–27.
Aula, L., Dhillon, J.  S., Omara, P., Wehmeyer, G.  B., Freeman, K.  W., & Raun, W.  R. (2019).
World sulfur use efficiency for cereal crops. Agronomy Journal, 111(5), 2485–2492. https://
doi.org/10.2134/agronj2019.02.0095
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting circular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Basu, S., Rabara, R.  C., & Negi, S. (2018). AMF: The future prospect for sustainable agricul-
ture. Physiological and Molecular Plant Pathology, 102, 36–45. https://doi.org/10.1016/j.
pmpp.2017.11.007
Bender, S.  F., Wagg, C., & van der Heijden, M.  G. A. (2016). An underground revolution:
Biodiversity and soil ecological engineering for agricultural sustainability. Trends in Ecology
and Evolution, 31(6), 440–452. https://doi.org/10.1016/j.tree.2016.02.016
Bertrand, M., Barot, S., Blouin, M., Whalen, J., de Oliveira, T., & Roger-Estrade, J. (2015).
Earthworm services for cropping systems. A review. Agronomy for Sustainable Development,
35(2), 553–567. https://doi.org/10.1007/s13593-­014-­0269-­7
Bhattacharyya, R., Tuti, M. D., Kundu, S., Bisht, J. K., & Bhatt, J. C. (2012). Conservation till-
age impacts on soil aggregation and carbon pools in a sandy clay loam soil of the Indian
Himalayas. Soil Science Society of America Journal, 76(2), 617–627. https://doi.org/10.2136/
sssaj2011.0320
Bisen, N., & Rahangdale, C. (2017). Crop residues management option for sustainable soil health
in rice-wheat system: A review. International Journal of Chemical Studies, 5, 1038–1042.
Briones, M. J. I., & Schmidt, O. (2017). Conventional tillage decreases the abundance and biomass
of earthworms and alters their community structure in a global meta-analysis. Global Change
Biology, 23(10), 4396–4419. https://doi.org/10.1111/gcb.13744
Bronick, C. J., & Lal, R. (2005). Soil structure and management: A review. Geoderma, 124(1–2),
3–22. https://doi.org/10.1016/j.geoderma.2004.03.005
Bruun, T. B., Elberling, B., de Neergaard, A., & Magid, J. (2015). Organic carbon dynamics in dif-
ferent soil types after conversion of forest to agriculture. Land Degradation and Development,
26(3), 272–283. https://doi.org/10.1002/ldr.2205
Bünemann, E. K., Bongiorno, G., Bai, Z., Creamer, R. E., De Deyn, G., de Goede, R., Fleskens,
L., Geissen, V., Kuyper, T.  W., Mäder, P., Pulleman, M., Sukkel, W., van Groenigen, J.  W.,
& Brussaard, L. (2018). Soil quality–A critical review. Soil Biology and Biochemistry, 120,
105–125. https://doi.org/10.1016/j.soilbio.2018.01.030
62 H. Bashir et al.

Butterly, C. R., Baldock, J. A., & Tang, C. (2013). The contribution of crop residues to changes
in soil pH under field conditions. Plant and Soil, 366(1–2), 185–198. https://doi.org/10.1007/
s11104-­012-­1422-­1
Celik, I., Günal, H., Acar, M., Gök, M., Bereket Barut, Z., & Pamiralan, H. (2017). Long-term
tillage and residue management effect on soil compaction and nitrate leaching in a Typic
Haploxerert soil. International Journal of Plant Production, 11, 131–149.
Chen, Z., Ti, J.-s., & Chen, F. (2017). Soil aggregates response to tillage and residue management
in a double paddy rice soil of the Southern China. Nutrient Cycling in Agroecosystems, 109(2),
103–114. https://doi.org/10.1007/s10705-­017-­9864-­8
Chowdhury, S., Farrell, M., Butler, G., & Bolan, N. (2015). Assessing the effect of crop residue
removal on soil organic carbon storage and microbial activity in a no-till cropping system. Soil
Use and Management, 31(4), 450–460. https://doi.org/10.1111/sum.12215
Dendooven, L., Patiño-Zúñiga, L., Verhulst, N., Luna-Guido, M., Marsch, R., & Govaerts,
B. (2012). Global warming potential of agricultural systems with contrasting tillage and residue
management in the central highlands of Mexico. Agriculture, Ecosystems and Environment,
152, 50–58. https://doi.org/10.1016/j.agee.2012.02.010
Deryng, D., Conway, D., Ramankutty, N., Price, J., & Warren, R. (2014). Global crop yield
response to extreme heat stress under multiple climate change futures. Environmental Research
Letters, 9(3), 034011. https://doi.org/10.1088/1748-­9326/9/3/034011
Dheri, G. S., & Nazir, G. (2021). A review on carbon pools and sequestration as influenced by
long-term management practices in a rice–wheat cropping system. Carbon Management,
12(5), 559–580. https://doi.org/10.1080/17583004.2021.1976674
Dias, T., Dukes, A., & Antunes, P.  M. (2015). Accounting for soil biotic effects on soil health
and crop productivity in the design of crop rotations. Journal of the Science of Food and
Agriculture, 95(3), 447–454. https://doi.org/10.1002/jsfa.6565
Din, W. M. U., Hussain, M. M., & Farooqi, Z. U. R. (2022). Climate change-induced aggravations
in microbial populations and processes: Constraints and remediations. In Climate change and
microbes (pp. 25–49). Apple Academic Press.
Eswaran, H., Lal, R., & Reich, P. (2019). Land degradation: An overview. In Response to land
degradation (pp. 20–35). CRC Press.
Farid Eltom, A. E., Ding, W., Ding, Q., Tagar, A. A., Talha, Z., & Gamareldawla. (2015). Field
investigation of a trash-board, tillage depth and low speed effect on the displacement and burial
of straw. CATENA, 133, 385–393. https://doi.org/10.1016/j.catena.2015.05.025
Farooqi, Z.  U. R., Hameed, M.  A., Mohy-Ud-Din, W., Ali, M.  H., Qadir, A., & Hussain,
M. M. (2021). Global climate change and microbial ecology: Current scenario and manage-
ment. In Microbiological activity for soil and plant health management (pp. 285–313). Springer.
Farooqi, Z. U. R., Din, W. M. U., & Hussain, M. M. (2022). Microbial responses under climate
change scenarios: Adaptation and mitigations. In Climate change and microbial diversity
(pp. 1–20). Apple Academic Press.
Fasinmirin, J.  T., & Reichert, J.  M. (2011). Conservation tillage for cassava (Manihot escu-
lenta Crantz) production in the tropics. Soil and Tillage Research, 113(1), 1–10. https://doi.
org/10.1016/j.still.2011.01.008
Feng, J., He, K., Zhang, Q., Han, M., & Zhu, B. (2022). Changes in plant inputs alter soil carbon
and microbial communities in forest ecosystems. Global Change Biology, 28(10), 3426–3440.
https://doi.org/10.1111/gcb.16107
Ginni, G., Kavitha, S., Yukesh Kannah, R., Bhatia, S. K., Adish Kumar, S., Rajkumar, M., Kumar,
G., Pugazhendhi, A., Chi, N.  T. L., & Rajesh Banu, J. (2021). Valorization of agricultural
residues: Different biorefinery routes. Journal of Environmental Chemical Engineering, 9(4),
105435. https://doi.org/10.1016/j.jece.2021.105435
Grzyb, A., Wolna-Maruwka, A., & Niewiadomska, A. (2020). Environmental factors affecting the
mineralization of crop residues. Agronomy, 10(12). https://doi.org/10.3390/agronomy10121951
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 63

Hameed, M. A., Farooqi, Z. U. R., Hussain, M. M., & Ayub, M. A. (2021a). PGPR-assisted biore-
mediation and plant growth: A sustainable approach for crop production using polluted soils. In
Plant growth regulators: Signalling under stress conditions (p. 403). Springer.
Hameed, M.  A., Farooqi, Z.  U. R., Younas, F., Din, W.  M. U., Hussain, M.  M., & Khilji,
S.  A. (2021b). Microplastic pollution: Sources, fate, impacts and research gaps. Quality of
Life, 20, 39–50.
Havlicek, E., & Mitchell, E.  A. (2014). Soils supporting biodiversity. In Interactions in soil:
Promoting plant growth (pp. 27–58). Springer.
Heikkinen, J., Ketoja, E., Seppänen, L., Luostarinen, S., Fritze, H., Pennanen, T., Peltoniemi, K.,
Velmala, S., Hanajik, P., & Regina, K. (2021). Chemical composition controls the decomposi-
tion of organic amendments and influences the microbial community structure in agricultural
soils. Carbon Management, 12(4), 359–376. https://doi.org/10.1080/17583004.2021.1947386
Heil, J., Vereecken, H., & Brüggemann, N. (2016). A review of chemical reactions of nitrifica-
tion intermediates and their role in nitrogen cycling and nitrogen trace gas formation in soil.
European Journal of Soil Science, 67(1), 23–39. https://doi.org/10.1111/ejss.12306
Hijbeek, R., van Ittersum, M. K., ten Berge, H. F. M., Gort, G., Spiegel, H., & Whitmore, A. P. (2017).
Do organic inputs matter–a meta-analysis of additional yield effects for arable crops in Europe.
Plant and Soil, 411(1–2), 293–303. https://doi.org/10.1007/s11104-­016-­3031-­x
Huera-Lucero, T., Labrador-Moreno, J., Blanco-Salas, J., & Ruiz-Téllez, T. (2020). A framework
to incorporate biological soil quality indicators into assessing the sustainability of territories in
the Ecuadorian Amazon. Sustainability, 12(7), 3007. https://doi.org/10.3390/su12073007
Hussain, M., Farooqi, Z., Mohy-Ud-Din, W., Younas, F., Shahzad, M., Ghani, M., Ayub, M., &
Qadeer, A. (2021a). Application of Bioremediation as sustainable approach to remediate heavy
metal and pesticide polluted environments. Plant and Environment, 1, 62–92.
Hussain, M. M., Farooqi, Z. U. R., Latif, J., Mubarak, M. U., & Younas, F. (2021b). Bioremediation:
A Green Solution to avoid Pollution of the Environment. In Soil bioremediation: An approach
towards sustainable technology (pp. 15–40). Wiley-Blackwell.
Hussain, M. M., Farooqi, Z. U. R., Rasheed, F., & Din, W. M. U. (2021c). Role of microorganisms
as climate engineers: Mitigation and adaptations to climate change. In Climate change and
microbes (pp. 1–24). Apple Academic Press.
Hussain, M. M., Mohy-Ud-Din, W., Younas, F., Niazi, N. K., Bibi, I., Yang, X., Rasheed, F., &
Farooqi, Z.  U. R. (2022). Biochar: A game changer for sustainable agriculture. Journal of
Sustainable Agriculture, 143–157. Springer.
Iqbal, M., Ul-Hassan, A., & van Es, H. M. (2011). Influence of residue management and tillage
systems on carbon sequestration and nitrogen, phosphorus, and potassium dynamics of soil
and plant and wheat production in semi-arid region. Communications in Soil Science and Plant
Analysis, 42(5), 528–547. https://doi.org/10.1080/00103624.2011.546929
Johan, P.  D., Ahmed, O.  H., Omar, L., & Hasbullah, N.  A. (2021). Phosphorus transformation
in soils following co-application of charcoal and wood ash. Agronomy, 11(10). https://doi.
org/10.3390/agronomy11102010
Kareem, A., Farooqi, Z.  U. R., Kalsom, A., Mohy-Ud-Din, W., Hussain, M.  M., Raza, M., &
Khursheed, M. M. (2022). Organic farming for sustainable soil use, management, food pro-
duction and climate change mitigation. Journal of Sustainable Agriculture, 39–59. Springer.
Kaur, G., Singh, G., Motavalli, P.  P., Nelson, K.  A., Orlowski, J.  M., & Golden, B.  R. (2020).
Impacts and management strategies for crop production in waterlogged or flooded soils: A
review. Agronomy Journal, 112(3), 1475–1501. https://doi.org/10.1002/agj2.20093
Lehman, R. M., Cambardella, C. A., Stott, D. E., Acosta-Martinez, V., Manter, D. K., Buyer, J. S.,
Maul, J. E., Smith, J. L., Collins, H. P., Halvorson, J. J., Kremer, R., Lundgren, J., Ducey, T.,
Jin, V., & Karlen, D. (2015). Understanding and enhancing soil biological health: The solu-
tion for reversing soil degradation. Sustainability, 7(1), 988–1027. https://doi.org/10.3390/
su7010988
64 H. Bashir et al.

Li, C.-f., Yue, L.-x., Kou, Z.-k., Zhang, Z.-s., Wang, J.-p., & Cao, C.-g. (2012). Short-term effects
of conservation management practices on soil labile organic carbon fractions under a rape–
rice rotation in central China. Soil and Tillage Research, 119, 31–37. https://doi.org/10.1016/j.
still.2011.12.005
Li, J., Wu, X., Gebremikael, M. T., Wu, H., Cai, D., Wang, B., Li, B., Zhang, J., Li, Y., & Xi,
J. (2018). Response of soil organic carbon fractions, microbial community composition and
carbon mineralization to high-input fertilizer practices under an intensive agricultural system.
PLoS One, 13(4), e0195144. https://doi.org/10.1371/journal.pone.0195144
Li, Y., Xu, J., Hu, J., Zhang, T., Wu, X., & Yang, Y. (2022). Arbuscular mycorrhizal fungi and
glomalin play a crucial role in soil aggregate stability in Pb-contaminated soil. International
Journal of Environmental Research and Public Health, 19(9), 5029. https://doi.org/10.3390/
ijerph19095029
Luo, Y., Iqbal, A., He, L., Zhao, Q., Wei, S., Ali, I., Ullah, S., Yan, B., & Jiang, L. (2020). Long-­
term no-tillage and straw retention management enhances soil bacterial community diver-
sity and soil properties in Southern China. Agronomy, 10(9), 1233. https://doi.org/10.3390/
agronomy10091233
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In S. A. Bandh (Ed.),
Climate change: The social and scientific construct. Springer Nature Publishing. https://link.
springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Menta, C. (2012). Soil fauna diversity-function, soil degradation, biological indices, soil restora-
tion. In Biodiversity conservation and utilization in a diverse world (pp. 59–94). InTech.
Mert, A., Celik, I., & Günal, H. (2018). Effects of long-term tillage systems on aggregate-­
associated organic carbon in the eastern Mediterranean region of Turkey. Eurasian Journal of
Soil Science, 7, 51–58.
Minas, A. M., Mander, S., & McLachlan, C. (2020). How can we engage farmers in bioenergy
development? Building a social innovation strategy for rice straw bioenergy in The Philippines
and Vietnam. Energy Research and Social Science, 70, 101717. https://doi.org/10.1016/j.
erss.2020.101717
Mirzaei, M., Gorji Anari, M., Razavy-Toosi, E., Asadi, H., Moghiseh, E., Saronjic, N., & Rodrigo-­
Comino, J. (2021). Preliminary effects of crop residue management on soil quality and crop
production under different soil management regimes in corn-wheat rotation systems. Agronomy,
11(2), 302. https://doi.org/10.3390/agronomy11020302
Mischler, J., Johnson, P. T., McKenzie, V. J., & Townsend, A. R. (2016). Parasite infection alters
nitrogen cycling at the ecosystem scale. Journal of Animal Ecology, 85(3), 817–828. https://
doi.org/10.1111/1365-­2656.12505
Mitchell, E., Scheer, C., Rowlings, D., Conant, R. T., Cotrufo, M. F., & Grace, P. (2018). Amount
and incorporation of plant residue inputs modify residue stabilisation dynamics in soil
organic matter fractions. Agriculture, Ecosystems and Environment, 256, 82–91. https://doi.
org/10.1016/j.agee.2017.12.006
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Navarro-Pedreño, J., Almendro-Candel, M.  B., & Zorpas, A.  A. (2021). The increase of soil
organic matter reduces global warming, myth or reality? Sci, 3(1), 18. https://doi.org/10.3390/
sci3010018
Naveen Kumar, B. T., & Babalad, H. B. (2018). Soil organic carbon, carbon sequestration, soil
microbial biomass carbon and nitrogen and soil enzymatic activity as influenced by conser-
vation agriculture in pigeonpea and soybean intercropping system. International Journal
of Current Microbiology and Applied Sciences, 7(3), 323–333. https://doi.org/10.20546/
ijcmas.2018.703.038
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 65

Nyakudya, I.  W., & Stroosnijder, L. (2015). Conservation tillage of rainfed maize in semi-arid
Zimbabwe: A review. Soil and Tillage Research, 145, 184–197. https://doi.org/10.1016/j.
still.2014.09.003
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Paterson, E., & Sim, A. (2013). Soil-specific response functions of organic matter mineraliza-
tion to the availability of labile carbon. Global Change Biology, 19(5), 1562–1571. https://doi.
org/10.1111/gcb.12140
Powlson, D. S., Whitmore, A. P., & Goulding, K. W. T. (2011). Soil carbon sequestration to miti-
gate climate change: A critical re-examination to identify the true and the false. European
Journal of Soil Science, 62(1), 42–55. https://doi.org/10.1111/j.1365-­2389.2010.01342.x
Prasad, J. V. N. S., Rao, C. S., Srinivas, K., Jyothi, C. N., Venkateswarlu, B., Ramachandrappa,
B. K., Dhanapal, G. N., Ravichandra, K., & Mishra, P. K. (2016). Effect of ten years of reduced
tillage and recycling of organic matter on crop yields, soil organic carbon and its fractions in
alfisols of semi arid tropics of southern India. Soil and Tillage Research, 156, 131–139. https://
doi.org/10.1016/j.still.2015.10.013
Qadir, A., Hameed, M., Bin Zafar, M., Farooqi, Z., Younas, F., Hussain, M., & Mohy-Ud-Din,
W. (2021). Phytoremediation of inorganic pollutants: An eco-friendly approach, its types and
mechanisms. Plant and Environment, 1, 110–129.
Sapkota, T. B. (2012). Conservation tillage impact on soil aggregation, organic matter turnover and
biodiversity. In Organic fertilisation, soil quality and human health (pp. 141–160). Springer.
Sidhu, H. S., Manpreet-Singh, Humphreys, E., Yadvinder-Singh, Balwinder-Singh, Dhillon, S. S.,
Blackwell, J., Bector, V., Malkeet-Singh, & Sarbjeet-Singh. (2007). The happy seeder enables
direct drilling of wheat into rice stubble. Australian Journal of Experimental Agriculture,
47(7), 844–854. https://doi.org/10.1071/EA06225
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. https://doi.org/10.16943/ptinsa/2014/v80i1/55089
Sithole, N. J., Magwaza, L. S., & Mafongoya, P. L. (2016). Conservation agriculture and its impact
on soil quality and maize yield: A South African perspective. Soil and Tillage Research, 162,
55–67. https://doi.org/10.1016/j.still.2016.04.014
Somasundaram, J., Sinha, N. K., Dalal, R. C., Lal, R., Mohanty, M., Naorem, A. K., Hati, K. M.,
Chaudhary, R.  S., Biswas, A.  K., Patra, 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/10.1080/07352689.2020.1782069
Sombrero, A., & De Benito, A. (2010). Carbon accumulation in soil. Ten-year study of conser-
vation tillage and crop rotation in a semi-arid area of Castile-Leon, Spain. Soil and Tillage
Research, 107(2), 64–70. https://doi.org/10.1016/j.still.2010.02.009
Thangarajan, R., Bolan, N. S., Tian, G., Naidu, R., & Kunhikrishnan, A. (2013). Role of organic
amendment application on greenhouse gas emission from soil. Science of the Total Environment,
465, 72–96. https://doi.org/10.1016/j.scitotenv.2013.01.031
Tiwari, S., Singh, C., Boudh, S., Rai, P. K., Gupta, V. K., & Singh, J. S. (2019). Land use change:
A key ecological disturbance declines soil microbial biomass in dry tropical uplands. Journal
of Environmental Management, 242, 1–10. https://doi.org/10.1016/j.jenvman.2019.04.052
Turmel, M.-S., Speratti, A., Baudron, F., Verhulst, N., & Govaerts, B. (2015). Crop residue man-
agement and soil health: A systems analysis. Agricultural Systems, 134, 6–16. https://doi.
org/10.1016/j.agsy.2014.05.009
Upadhyay, S. K., Singh, J. S., Saxena, A. K., & Singh, D. P. (2012). Impact of PGPR inocula-
tion on growth and antioxidant status of wheat under saline conditions. Plant Biology, 14(4),
605–611. https://doi.org/10.1111/j.1438-­8677.2011.00533.x
66 H. Bashir et al.

Urra, J., Alkorta, I., & Garbisu, C. (2019). Potential benefits and risks for soil health derived from
the use of organic amendments in agriculture. Agronomy, 9(9), 542. https://doi.org/10.3390/
agronomy9090542
Valkama, E., Kunypiyaeva, G., Zhapayev, R., Karabayev, M., Zhusupbekov, E., Perego, A.,
Schillaci, C., Sacco, D., Moretti, B., Grignani, C., & Acutis, M. (2020). Can conservation agri-
culture increase soil carbon sequestration? A modelling approach. Geoderma, 369, 114298.
https://doi.org/10.1016/j.geoderma.2020.114298
Vanhie, M., Deen, W., Lauzon, J. D., & Hooker, D. C. (2015). Effect of increasing levels of maize
(Zea mays L.) residue on no-till soybean (Glycine max Merr.) in Northern production regions:
A review. Soil and Tillage Research, 150, 201–210. https://doi.org/10.1016/j.still.2015.01.011
Verhulst, N., Carrillo-García, A., Moeller, C., Trethowan, R., Sayre, K. D., & Govaerts, B. (2011a).
Conservation agriculture for wheat-based cropping systems under gravity irrigation: Increasing
resilience through improved soil quality. Plant and Soil, 340(1–2), 467–479. https://doi.
org/10.1007/s11104-­010-­0620-­y
Verhulst, N., Kienle, F., Sayre, K. D., Deckers, J., Raes, D., Limon-Ortega, A., Tijerina-Chavez,
L., & Govaerts, B. (2011b). Soil quality as affected by tillage-residue management in a
wheat-maize irrigated bed planting system. Plant and Soil, 340(1–2), 453–466. https://doi.
org/10.1007/s11104-­010-­0618-­5
Victoria, R., Banwart, S., Black, H., Ingram, J., Joosten, H., Milne, E., Noellemeyer, E., &
Baskin, Y. (2012). The benefits of soil carbon. Foresight chapter in UNEP [Yearbook], 2012
(pp. 19–33).
Walsh, O. S., & Belmont, K. M. (2015). Improving nitrogen-use efficiency in Idaho crop produc-
tion. University of Idaho Extension.
Wander, M. M., Cihacek, L. J., Coyne, M., Drijber, R. A., Grossman, J. M., Gutknecht, J. L. M.,
Horwath, W. R., Jagadamma, S., Olk, D. C., Ruark, M., Snapp, S. S., Tiemann, L. K., Weil, R.,
& Turco, R. F. (2019). Developments in agricultural soil quality and health: Reflections by the
research committee on soil organic matter management. Frontiers in Environmental Science, 7,
109. https://doi.org/10.3389/fenvs.2019.00109
Wang, H., Wang, S., Yu, Q., Zhang, Y., Wang, R., Li, J., & Wang, X. (2020). No tillage increases
soil organic carbon storage and decreases carbon dioxide emission in the crop residue-­
returned farming system. Journal of Environmental Management, 261, 110261. https://doi.
org/10.1016/j.jenvman.2020.110261
Wei, L., Vosátka, M., Cai, B., Ding, J., Lu, C., Xu, J., Yan, W., Li, Y., & Liu, C. (2019). The role of
arbuscular mycorrhiza fungi in the decomposition of fresh residue and soil organic carbon: A
mini-review. Soil Science Society of America Journal, 83(3), 511–517. https://doi.org/10.2136/
sssaj2018.05.0205
West, T.  O., & Six, J. (2007). Considering the influence of sequestration duration and carbon
saturation on estimates of soil carbon capacity. Climatic Change, 80(1–2), 25–41. https://doi.
org/10.1007/s10584-­006-­9173-­8
Wijesinha-Bettoni, R., & Mouillé, B. (2019). The contribution of potatoes to global food security,
nutrition and healthy diets. American Journal of Potato Research, 96(2), 139–149. https://doi.
org/10.1007/s12230-­018-­09697-­1
Yang, X., Drury, C.  F., & Wander, M.  M. (2013). A wide view of no-tillage practices and soil
organic carbon sequestration. Acta Agriculturae Scandinavica, Section B  – Soil and Plant
Science, 63(6), 523–530. https://doi.org/10.1080/09064710.2013.816363
Younas, F., Mustafa, A., Farooqi, Z. U. R., Wang, X., Younas, S., Mohy-Ud-Din, W., Ashir Hameed,
M., Mohsin Abrar, M., Maitlo, A. A., Noreen, S., & Hussain, M. M. (2021). Current and emerg-
ing adsorbent technologies for wastewater treatment: Trends, limitations, and environmental
implications. Water, 13(2), 215. https://doi.org/10.3390/w13020215
Zhang, H., Zhang, Y., Yan, C., Liu, E., & Chen, B. (2016). Soil nitrogen and its fractions between
long-term conventional and no-tillage systems with straw retention in dryland farming in north-
ern China. Geoderma, 269, 138–144. https://doi.org/10.1016/j.geoderma.2016.02.001
4  Crop Residue Incorporation to Enhance Soil Health in the Rice–Wheat System 67

Zhang, Y., Li, X., Gregorich, E. G., McLaughlin, N. B., Zhang, X., Guo, Y., Liang, A., Fan, R.,
& Sun, B. (2018). No-tillage with continuous maize cropping enhances soil aggregation and
organic carbon storage in Northeast China. Geoderma, 330, 204–211. https://doi.org/10.1016/j.
geoderma.2018.05.037
Zhou, X., Wu, H., Li, G., & Chen, C. (2016). Short-term contributions of cover crop surface
residue return to soil carbon and nitrogen contents in temperate Australia. Environmental
Science and Pollution Research International, 23(22), 23175–23183. https://doi.org/10.1007/
s11356-­016-­7549-­5
Zuo, L., Zhang, Z., Carlson, K.  M., MacDonald, G.  K., Brauman, K.  A., Liu, Y., Zhang, W.,
Zhang, H., Wu, W., Zhao, X., Wang, X., Liu, B., Yi, L., Wen, Q., Liu, F., Xu, J., Hu, S., Sun,
F., Gerber, J. S., & West, P. C. (2018). Progress towards sustainable intensification in China
challenged by land-use change. Nature Sustainability, 1(6), 304–313. https://doi.org/10.1038/
s41893-­018-­0076-­2
Chapter 5
Promoting Energy Crops to Replace Fossil
Fuel Use

Muhammad Irfan, Liu Xianhua, Asia Shauket, Muhammad Jafir,


Adeel Ahmad, Samina Jam Nazeer Ahmad, and Jam Nazeer Ahmad

Abstract  Biomass energy development is a critical component in resolving the


energy issue and mitigating the effects of global warming. Renewable energy
sources such as biomass can be described as carbon neutral, as they emit only car-
bon that has been trapped in the plant growth cycle. Biofuel is the only liquid fuel
that can be used as an alternative to fossil fuels for transportation. Biofuels have the
potential to address future global energy demand. Energy crops referred to as energy
farming is another option for agriculture on agricultural lands. In terms of cost and
environmental benefits, energy crops for biofuel generation have been studied
extensively. Energy is now obtained from commodities including seed crops and
some C4 crops, like sweet sorghum, switchgrass, and miscanthus, can grow on
infertile terrain and produce substantial amounts of biomass. Energy crops now
make up a small percentage of the total energy provided by biomass, but this is
expected to change in the next decades. This chapter includes (1) global energy clas-
sification, (2) biofuel-producing energy crops, (3) potential contribution of energy
crops to replace fossil fuel, (4) conversion technologies and (5) prospects of
energy crops.

M. Irfan (*) · L. Xianhua


School of Environmental Science and Engineering, Tianjin University, Tianjin, PR China
e-mail: irfan@tju.edu.cn
A. Shauket · S. J. N. Ahmad
Department of Botany, University of Agriculture Faisalabad, Faisalabad, Pakistan
M. Jafir · J. N. Ahmad
Department of Entomology, University of Agriculture Faisalabad, Faisalabad, Pakistan
A. Ahmad
Institute of Soil and Environmental Science, University of Agriculture Faisalabad, Faisalabad,
Pakistan

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


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_5
70 M. Irfan et al.

1 Introduction

In many debates, climate change and global warming are emphasized as fundamen-
tal threats, particularly in this age of globalization where country borders and spatial
heterogeneity are being reduced (Usman et  al., 2021). It is challenging to put a
precise numerical value on the magnitude of fossil fuel subsidies. In 2015, it was
predicted that fossil fuel subsidies amounted to $4.7 trillion (6.3%) of the world’s
real GDP, and in 2017, it was projected that they would rise to $5.2 trillion (6.5%)
(Coady et  al., 2019).). In 2016, the G7 leaders issued a call to action, urging all
governments to end fossil fuel subsidies by 2025. There are many facets of fossil
fuel subsidies that have not been sufficiently studied in the literature, despite the
growing quantity of these subsidies around the world and the growing worry about
the increase in these subsidies. To name a few of these traits, subsidies for fossil
fuels are controversial because of the negative effects they have on the economy and
the environment, which have not been adequately addressed (Malla et  al., 2022;
Bandh et al., 2021, 2023; Mushtaq et al., 2020). The extent of ecological damage
was calculated by using the ecological footprint. The current global population of
7.4 billion people is expected to grow to 10 billion in 2055, which could put a strain
on the planet’s limited resources. Alternatively, human yearly use of biomass mate-
rials is estimated at 72 gigatons, rising to 100 gigatons by 2030 (Usmani et al., 2020).
To battle global climate concerns and reduce reliance on fossil fuels, the expan-
sion of the global economy is currently up against a unique challenge in many sec-
tors, including energy, food, and agriculture. There is now a global race to develop
commercially viable renewable biofuels from lignocellulosic biomass as a response
to rising fuel prices and the release of hazardous gases from the combustion of fossil
fuels (Islam et  al., 2020). The increased consumption of these nonrenewable
resources, however, poses a critical risk to the environment. Using a biorefinery
process to convert lignocellulose biomass into marketable industrial bioproducts
like renewable energy is a new but exciting area of study (Abraham et al., 2020).
Most lignocellulosic biomass comes from forest waste, farm waste, and energy
crops that are grown specifically for energy, organic municipal solid waste, and
industrial waste (wood, paper, and pulp). To minimize the worldwide environmental
impact of energy production and consumption, several renewable energy sources
(RES) can be utilized, including the conversion of biomass into usable energy (Roy
et al., 2021; Parray et al., 2022; Bandh et al., 2022; Bandh, 2022a, b).

2 Global Energy Classification

The global energy classification is based on nonrenewable and renewable energy


sources which can be utilized for power generation (Fig. 5.1). The nonrenewable
energy sources are crude oil, coal, natural gas, and nuclear. Nonrenewable energy
sources are those whose economic worth could not be recovered by natural means
5  Promoting Energy Crops to Replace Fossil Fuel Use 71

Fig. 5.1  Flowchart of global energy classification. (Elavarasan, 2019)

at the same level of use. Nonrenewable sources form over billions of years. They are
not sustainable (Elavarasan, 2019).
Renewable energy sources are currently replacing nonrenewable energy sources
since they are abundant and environmentally friendly. Renewable energy sources
include solar, wind, geothermal, hydrothermal, and biomass. Countries are cur-
rently focusing on renewable energy to lessen the greenhouse gas effect. Additionally,
hybrid systems consisting of two system combinations (renewable-nonrenewable)
are created to improve output (Guo et al., 2018).

3 Biomass Energy

Biomass-based energy is the most promising option for meeting the needs and
ensuring a steady supply of eco-friendly energy and fuel because of its high poten-
tial and wide applicability. There are numerous approaches that can be taken to
effectively utilize biomass resources. One of these is through developing and imple-
menting new biomass technologies that allow for more effective biomass produc-
tion and utilization. Agricultural biomass includes both the edible parts (oil and
simple carbohydrates) and the nonedible parts of crops including stems, leaves,
straws, etc. In the biomass context, agriculture-based biomass is a particularly large
group (Chandra et al., 2012). Numerous crops can produce starch, cellulose, and oil,
while some other crops produce biogas which can be utilized for heat and electricity
generation. Landfills produce methane which can be utilized as biofuels that com-
prise biodiesel, methanol, ethanol, and their derivatives (Demirbaş, 2001). These
are types of crops which can be used to create alternative energy sources. Significant
incentives exist now to encourage renewable fuel utilization in the transportation
72 M. Irfan et al.

sector across the globe. Several countries’ energy policies, climate policies, and
agricultural policies reportedly offer incentives to promote the continued produc-
tion and use of biofuels in the future. Carbon dioxide (CO2) levels in the air are
going up because fossil fuels are being burned (DGS, 2005; Kamm et al., 2006).
This is a big part of what is causing global warming. Concern over climate
change on a global scale, which is mostly brought on by the combustion of fossil
fuels, is one of the primary motivating factors behind the expansion of the biofuel
industry around the world. There is much evidence that greenhouse gas emissions
are to blame for the accelerated rate of global warming. The greenhouse effect is
exacerbated by carbon dioxide (Rutz & Janssen, 2007). Biofuels can be made from
a variety of renewable energy sources, but plant biomass has been seen as particu-
larly promising for decades. This is due to its several positive qualities, including its
low cost, negligible impact on the environment, and ample supply. About 14% of
the world’s energy needs are met by biomass, making it a major resource for the
economy. Additionally, the use of plant biomass has the potential to preserve and
improve both ecological and social sustainability, as well as contribute to the stabi-
lization of the incomes of farmers (Zhao et al., 2009). The world’s economy has
been growing at a rapid rate recently due to which the demand for energy is increas-
ing. The swift growth of major economies around the world, such as China, the
United States, and Europe, is sure to increase the need for energy. It is for this rea-
son that the dissemination of technologies that might fundamentally cut down on
carbon emissions is difficult (Li & Zhang, 2017). One of the essential components
of business expansion, technological advancement, and economic competitiveness
in modern economies is the energy industry (Muntean et al., 2018). An economy
that is focused on sustainability will have as some of its priorities the securing of
energy supplies and the protection of the environment. Initiatives based on collabo-
ration may have a beneficial effect on the environment (Cherry & Pidgeon, 2018).
The demand for fossil fuels will be able to fall because of the development of renew-
able energy sources (RES). Furthermore, compared to 2005, the number of green-
house gases emitted will be reduced by one-tenth (Searchinger et al., 2018).

3.1 Bioenergy Market

The growth of a sustainable bioenergy market that is supported by bioresources is


required to accomplish the goals of protecting the climate and maximizing the effi-
ciency of resources. To equalize the support to produce biofuels, there is a need, as
stated by (Zieliński et al., 2019) to look for nonfood biomass. Because of their role
in maintaining energy stability, renewable energy sources should be prioritized for
research and development to benefit the economy. Their contribution to the savings
associated with the import of fossil fuels in 2015 was 15 million euros, and it is
projected to be 58 million euros in the year 2030 (Study on the role of technical help
in the production of the report on renewable energy for 2016). The use of renewable
sources of energy in the EU carries with it significant implications for the region’s
5  Promoting Energy Crops to Replace Fossil Fuel Use 73

standing in the race for technological advancement. The European Union (EU)
owns 30% of all patents related to energy sources worldwide. However, a significant
number of businesses based in the EU struggle to accomplish their sustainable
goals. To match the incentives and income mechanisms necessary to drive sustain-
able solutions, therefore, innovation at the level of the business model is essential
(Geissdoerfer et al., 2018; Rashid et al., 2013). It is estimated that most of the world
population will prefer to live in urban areas until 2050 (Desa, 2017), which presents
a huge problem for policymakers attempting to promote sustainability (Petit-Boix
& Leipold, 2018). Greenhouse gas emissions are lower in rural areas, but farmers
can afford to make the necessary investments in their fields, especially in cutting-­
edge equipment that will allow them to boost production on a larger scale. Although
the price of agricultural inputs can fluctuate, the demand for food is fixed, so these
efforts will not lead to higher revenue, and farmers are being pushed to invest in
more efficient technology because of this trend where the marginal revenue from
agricultural output is falling (Czyżewski et al., 2019).

4 Agricultural Biofuels

The use of renewable energy sources, such as biomass, to generate power is one
approach to lessening the global environmental impact of energy production and
use (Owusu & Asumadu-Sarkodie, 2016). Biomass is used to make energy in five
different ways: growing plants for sugar, starch, cellulose, and oil, burning waste,
using anaerobic digesters to make biogas that can be used to make heat or electric-
ity, making gas from landfills, and making biofuels like methanol, biodiesel, etha-
nol, and their derivatives (Grangeiro et al., 2019). Concern over climate change on
a global scale, which is mostly brought on by the combustion of fossil fuels, is one
of the primary motivating factors behind the expansion of the biofuel industry
around the world. There is much evidence that greenhouse gas emissions are to
blame for the accelerated rate of global warming. The greenhouse effect is intensi-
fied by carbon dioxide (Prasad et al., 2020). Biofuels can be made from a variety of
renewable energy sources, but plant biomass has been seen as particularly promis-
ing for decades. This is due to its several positive qualities, including its low cost,
negligible impact on the environment, and ample supply. About 14% of the world’s
energy needs are met by biomass, making it a major resource for the economy. Plant
biomass has the potential to preserve and improve both ecological and social sus-
tainability, as well as to contribute to the stabilization of the incomes of farmers
(Bórawski et al., 2019). The use of oils derived from soybeans, canola, corn, rape-
seed, and palm oil is generally acknowledged to be among the basic ingredients for
biodiesel production. New plant oils, including those made from mustard seeds,
peanuts, sunflowers, and cotton seeds, are currently under study (Eryilmaz et al.,
2016). When making biodiesel, several different kinds of bio-lipids can be used.
Some other crops are hemp, palm oil, mustard, sunflower, and some types of algae
used as a replacement for vegetable oil that is not in use and fats obtained from
74 M. Irfan et al.

animals (Khatib & Yassine, 2019). Only the oils from peanuts, soybeans, palms,
sunflowers, safflowers, cottonseeds, and rapeseeds are investigated as possible
replacement fuels for diesel vehicles. This is out of more than 350 different oil-­
producing crops (Eryilmaz et al., 2016).

4.1 Sweet Sorghum, a Source of Ethanol Production

Annual C4 sweet sorghum is notable for its high photosynthetic efficiency. It is a


plant that produces a lot of biomasses and has a high glucose content. Sucrose (up
to 55%) and glucose (up to 3.2%) make up the bulk of their stalks (Diallo et al.,
2019). Cellulose accounts for 12.4%, whereas hemicellulose makes up 10.2% of its
composition. There are a lot of fermentable sugars in sweet sorghum biomass, so it
is commonly thought of as a great starting point for making fermentative hydrogen.
When compared to the other “new crops” that are now being studied as potential
energy and industrial raw material sources, sweet sorghum appears to have the most
promise (Diallo et al., 2019). Today, ethanol and methane are the two most well-­
known by-products of sweet sorghum fermentation. Assuming an energy yield of
26,500  kJ/kg, the cited research indicated that the energy obtained from ethanol
varied between 6500 and 8900 kJkg−1 of dry sorghum biomass and between 1400
and 2700 kJkg−1 from fresh sorghum biomass (Diallo et al., 2019).

Brassica napus (Rapeseed)


4.2 

Brassica oil crops, which include Brassica napus (rapeseed), is referred to as oil-
seed rape (Indian mustard). The second highest-yielding oil crop worldwide is rape-
seed. This crop is known as the energy crop that receives the greatest land in
cultivation in Europe and is grown for the production of fodder, biofuel, and food.
Rapeseed oil production made up 14% of oil crop production worldwide and was
among the top 14 agricultural products in Europe. The nonedible oil demand is ris-
ing around the globe as a result of the growing industrialization of developing
nations. Energy is a strategic national concern since petroleum supplies are running
out at the same time. Therefore, it is essential to produce clean and effective bioen-
ergy. For a number of reasons, rapeseed is the best crop for the manufacture of
biodiesel. This crop produces a large quantity of oil because it has a lot of cellulose.
Because of this, it is currently the most popular choice for making biodiesel. It may
be the only alternative that can be used directly as a liquid biofuel without any
preparation. Firstly, canola/rapeseed has a 40% oil content and can yield a lot of oil
per unit of space. Secondly, rapeseed biodiesel has a significantly lower cloud point
(a measure of the oil’s propensity to clog filters) and pour point (the temperature at
which the liquid turns semisolid and loses its flow properties). Canola biodiesel is
5  Promoting Energy Crops to Replace Fossil Fuel Use 75

therefore a better fuel for colder climates because it is converted into gels at a lower
temperature than biodiesel made from other feedstocks. Rapeseed is a significant
source of biodiesel’s raw materials in temperate climate regions like Europe, where
it makes up 50–70% of the fuel. Even if the technique for use is fully developed,
there are currently few consistent sources of raw materials available. Therefore,
there is a need to develop varieties that are suitable for marginal soils and setting up
raw material processing equipment is crucial for the growth of the rapeseed bio-
diesel in the world (Fu et al., 2016; Popp et al., 2016; Raman et al., 2019; Van Duren
et al., 2015).

4.3 Jatropha (Jatropha curcas)

There are now cultivations in the southeast, north, and northeast of Brazil where
jatropha (Jatropha curcas) is grown to extract its oil, which is then used in the pro-
duction of biodiesel fuel in the Philippines and Brazil. Similarly, hundreds of initia-
tives in poor nations promote jatropha oil as an accessible biofuel crop (Rozina
et al., 2017). The cake that is left over after pressing the seeds can be used as fertil-
izer, fuel for vehicles, or even animal feed. Also, digesters can use the whole seed,
including the oil, to make biogas. Jatropha curcas, also known as jatropha, and
Pongamia pinnata are the two types of raw materials that are now used in produc-
tion in India (Karanja). Because of the presence of toxins, neither of these plants’
oils should be consumed in any form. Both jatropha and Karanja trees produce oil-­
rich seeds, with jatropha seeds containing 40% and Karanja seeds containing 33%
(Demirbas et al., 2016).

4.4 Palm Oil (Elaeis guineensis)

Its scientific name is Elaeis guineensis, which is native to West Africa. It was ini-
tially discovered to grow naturally in that region, but it was later farmed as an agri-
cultural commodity once it was discovered there. Since palm oil comes from a
tropical perennial plant, it is best grown in lowlands that experience high levels of
humidity. As a result, palm oil might be easily cultivated in Malaysia. There is only
one trunk on the tree, and it does not have any branches. The tree’s height can reach
between 20 and 30 m (Wakil et al., 2015). Each solitary flower is relatively small
and consists of three sepals, three petals, and three stamens. The blooms are pro-
duced in dense clusters. The leaves have a structure known as pinnate, and their
length can range anywhere from 3 to 5 m.
76 M. Irfan et al.

4.5 Safflower (Carthamus tinctorius L.)

To make biodiesel, its seed oil undergoes chemical processing known as the trans-
esterification reaction, which takes place in an atmosphere containing methyl alco-
hol and sodium hydroxide (NaOH) (Suresh et  al., 2018). Safflower, which is a
member of the Composite family and can thrive in a variety of climates, is grown in
several different regions of the globe. It is an annual plant that resembles a thistle
and is herbaceous. The leaves often have many very long and sharp spines on them.
The plants can grow up to 150 cm tall and have globular flower heads that bloom in
July. The flowers are typically a beautiful yellow, orange, or red color and each
flower head on a branch typically has 15–20 seeds in it and can have anywhere from
one to five bloom heads (Aydın, 2016).

4.6 Soybean (Glycine max)

When it comes to protein and oil, it is indispensable. Major producers include China
(7%), Argentina (21%), Brazil (27%), and the United States (33%). Soybeans can
also be extracted for protein and oil. This makes them a significant food crop. While
the protein is mostly used for animal feed, it is also used in some food products. The
oil, on the other hand, is used in a wider variety of food products, animal feed prod-
ucts, and even some industrial uses. One of the primary ingredients in biodiesel is
soybean oil. Over 80% of Brazil’s biodiesel comes from soybeans. When compared
to the United States (74%), the European Union (16%), and Argentina (100%), it is
at the 100% level. Therefore, soy holds great potential as a feedstock for making
biodiesel. Soybean oil and protein are in high demand, making genetic breeding for
increased yield and quality a pressing concern (André Cremonez et al., 2015).

5 Conversion Technologies

The production of a single unit of bioenergy results in carbon dioxide emissions that
are 10–20 times fewer than those associated with the production of fossil fuels
(Gabrielli et al., 2020). Biochemical technology allows for the breakdown of cel-
lulose into sugars and glycerides, which can then be processed further in biorefiner-
ies to produce chemical intermediates, bioethanol, and biodiesel. While the energy
input and output ratios are not exactly small, they are also not significant. Many
times, fossil fuels are used in the creation of bioenergy carriers. However, this
source of energy production accounts for a negligible fraction of the whole.
Bioenergy forestry and agriculture systems typically have energy conversion ratios
of 1:25 and 1:50, respectively (Schipfer, 2017). There is a diverse selection of bio-
energy carriers, ranging from unprocessed agricultural waste to highly refined
5  Promoting Energy Crops to Replace Fossil Fuel Use 77

biofuels for transportation. Several factors, such as availability, location of consum-


ers, and the presence or absence of coproducts or benefits, can all have an impact on
the objectives of using biomass. Products derived from biomass can be put to a
broad variety of different uses (Datta et al., 2019).

5.1 Combustion and Cofiring

One of the most common ways in which biomass is transformed is by combustion.


It is possible that in the future, as more is learned about the fundamentals of com-
bustion performance and ash behavior, plant dependability and efficiency will
improve. Emissions and localized investment costs can be decreased by improving
our knowledge of combustion (Ilic et al., 2018). Coffering solid biomass particles
with coal, mixing diesel with biodiesel and gasoline with bioethanol, and installing
vehicles with flexible fuel engines are all straightforward examples of merging bio-
mass and fossil fuel technologies. Co-using biomass materials in coal-fired boiler
systems have advanced rapidly in recent years. Co-firing has been used with a vari-
ety of biomass materials and commercially relevant fuels such as lignites, anthra-
cites, bituminous and subbituminous coals, and petroleum coke. Herbaceous and
sappy plants, trees, dry and wet farm trash, and energy crops all fall into this cate-
gory (Nelson et al., 2018).

5.2 Gasification

Biomass gasification involves the transformation of organic material into a gas or


vapor and a solid. Syngas, in their gaseous form, can be converted into energy or
biofuels to their high heating capacity. Char is the solid phase, and it is made up of
the inert residue from the biomass and any organic material that was not changed
throughout the treatment process. This change happens when some of the carbon in
the material being fed is oxidized. This usually happens in the presence of air,
steam, oxygen, and carbon dioxide. Many strategies are being investigated to maxi-
mize the quantity of biomass utilized to generate energy. One such strategy is the
gasification of biomass. As the price of oil continues to skyrocket, so does public
concern over the environmental impact of burning fossil fuels, which in turn has
sped up research into gasification technology for biomass. Produced at temperatures
between 250 and 300  °C, syngas is a gaseous combination of hydrogen, carbon
monoxide, carbon dioxide, methane, and heavier hydrocarbons (such as tars) that
condense. In addition to syngas, lighter hydrocarbons like ethane and propane are
also generated. Besides inert gases like nitrogen, syngas can also contain noxious
ones like sulfuric (H2S) and chloride (HCl) acid. Their existence is conditional on
the biomass type being gasified and the conditions under which the process is being
carried out. There is a wide range in the LHV of syngas, from 4 to 13 MJ/Nm3,
78 M. Irfan et al.

which is dependent on the feedstock, gasification method, and operating circum-


stances (Liu & Ji, 2013; Qian et al., 2013; Wu et al., 2014).
The combined ash and unconverted organic component produce char, which is
rich in carbon. The amount of organic matter that remains after gasification is deter-
mined by the technology used and the operating conditions. On the other hand, the
quantity of ash produced is proportional to the type of biomass that was incinerated.
Depending on how much of the organic fraction has been transformed, the char’s
LHV can range from 25 to 30 MJ/kg (Molino et al., 2016). It is the oxidation of
biomass, which can occur in either an all-thermal or autothermal phase, that often
supplies the energy needed for the major reactions that occur during gasification,
which are endothermic. By engaging in a process known as incomplete combustion,
the gasifier in an autothermal process generates its heat. In the all-thermal process,
on the other hand, energy is brought in from the outside to do the gasification at the
right level (Molino et al., 2016).

5.3 Biogas Generation

Biogas is a sustainable fuel obtained from the anaerobic breakdown of different


biological-based feedstocks via cooperative metabolic activities of acidogenic,
hydrolytic, and methanogenic bacteria. Biogas contains approximately methane
(60%), CO2 (40%), and H2S (2000 ppm), which are the major impurities. Methane
capture during biogas production helps to reduce CH4 emissions, and it can also be
used as a sustainable energy source for all applications that require natural gas. In
2014 biogas production enhanced (from 0.28 Ej to 1.28 Ej) with a volume of 59
billion m3. Many countries throughout the world grow corn primarily for the pro-
duction of biogas (Nikkhah et al., 2020; Scarlat et al., 2018; Villadsen et al., 2019).

5.4 Bioethanol Production

Alternatives to gasoline that are renewable and environmentally friendly include


bioethanol. It is simple to add an oxygenated component for cleaner combustion to
gasoline. Bioethanol production consists of different steps including initial treat-
ment, fermentation, hydrolysis, recovery, and refinement. The use of bioethanol as
a fuel began during the worldwide fuel crisis in the 1970s, and as a result of its
widespread use across numerous industries, its production capacity increased from
a lesser extent of 1 billion L (1975) to 39 billion L (2006). The main source of bio-
ethanol production is corn, a crop with high sugar or starch content. Since it is
widely planted around the world, corn produced 817 million tons worldwide in
2009, more than both rice (678 million tons) and wheat combined (682 million tons)
(Nikkhah et al., 2020; Sirajunnisa & Surendhiran, 2016; Thangavelu et al., 2016).
5  Promoting Energy Crops to Replace Fossil Fuel Use 79

5.5 Biodiesel Production

Without requiring significant hardware modifications, biodiesel, a substitute fuel for


diesel, can be used in standard diesel engines. “Bio” suggests that it comes from a
natural, renewable source, while “diesel” indicates that it is used as a fuel for diesel-­
powered motors. All through the process of transesterification, biodiesel can be
made using oilseeds like canola, sunflower, peanut, and soybean. In several nations,
including Italy, Germany, Turkey, and France, it might be the best alternative fuel.
The world’s top biodiesel producers in 2014 are shown in Fig. 5.2. One of the amaz-
ing resources for the production of biodiesel is oilseeds. In this context, peanuts are
recognized as one of the primary oilseed sources for the manufacturing of biodiesel,
and the initial biofuel to run a diesel engine was made from peanuts. Renewability
and biodegradability, a higher flash point, and the lack of sulfur and aromatic com-
pounds are the benefits of biodiesel. However, the creation of feedstock involves the
use of several inputs, like diesel fuel and chemical fertilizers, which might increase
greenhouse gas emissions, whenever the origin of its production is oilseeds
(Eryilmaz et al., 2016; Zhang & Balasubramanian, 2016).

5.0 4.7
4.5
Total production (billion litters)

4.0
3.4 3.4
3.5 3.1
2.9
3.0

2.5
2.1
2.0

1.5 1.1 1.2


1.0 0.7 0.7 0.8 0.8
0.6
0.5 0.3

0.0
s y a e s da n
te zi
l
an na in si
a
nc nd nd um ai nd bi
a
St
a a hi t e a la ai
la
an
a i Sp la
Br rm C en on Fr er lg Po om
d e rg nd h Th C Be ol
ni
te G A I et C
U N

Countries

Fig. 5.2  The largest biodiesel-producing countries globally. (Hajjari et al., 2017)
80 M. Irfan et al.

Biogas
62112.56 km

Bioethanol
13768.51 km

Biodiesel

6229.44 km

Fig. 5.3  The maximum distance a car may go using different biofuels (an average fuel usage of
10 L/100 km was estimated). (Nikkhah et al., 2020)

5.6 Green Technology Comparison

The production of biodiesel (59,204.61  MJha−1), bioethanol (26,786.58  MJha−1),


and biogas (267,084.00  MJha−1) comes from different energy crop sources. The
findings showed unequivocally that using corn silage as a source of biogas produced
more net energy than the other two biofuel production methods. The range that a
typical car can cover while using different biofuels is shown in Fig. 5.3 (1 ha energy
crop utilization for biofuel), assuming a 10-L gasoline consumption rate per 100 km.
According to estimates, a car can go 62,000 km on biogas, 14,000 km on bioetha-
nol, and 6000 km on biodiesel. In comparison to bioethanol and biodiesel, it means
that biogas has the best potential for use as a transportation fuel. In some nations,
biogas is used as a clean-burning transportation fuel. The European Union has also
established a goal to boost the use of biofuels; more specifically, 10% of the fuels
used in the transport industry should be made from biofuels by 2020, and the per-
centage should continue to rise after that year. Moreover, Iran has promised to
reduce its GHG emissions following the Paris Agreement. Therefore, the utilization
of upgraded biogas as a fossil fuel replacement in the transport industry could con-
tribute to the reduction of greenhouse gas emissions. Since 2002, the use of biogas
in urban transportation alone in Sweden has reduced the country’s annual carbon
dioxide emissions by 9000 metric tons (Ahmad et al., 2017; Nikkhah et al., 2020).

6 Carbon Offsets

When bioenergy is used in place of fossil fuels, carbon dioxide emissions are cut
immediately. Therefore, the benefits of carbon mitigation measures can be maxi-
mized by a combination of energy crop production including carbon sinks and
5  Promoting Energy Crops to Replace Fossil Fuel Use 81

offset credits. To accomplish this goal and enhance the average carbon stock in the
soil while simultaneously producing a source of biomass, energy crops can be
planted in areas that were formerly used for agriculture or pasture. Reusing the car-
bon stored in biomass for energy production helps solve the essential problem of
long-term preservation of the biotic carbon stores, as in the case of a permanent
forest. There is some evidence that growing perennial energy crops can boost soil
carbon levels. However, this idea is still in its infancy and needs more research,
including in-depth life cycle analyses for certain crops grown in different locations.
Over the next few decades, the global percentage of bioenergy may decline if laws
are not put in place to encourage novel bioenergy technology and sustainable bio-
mass production plans. Water use, biodiversity, and societal and economic concerns
might all take a major hit without the correct parameters in place. Using integrated
assessment models over the long run, the feasibility and cost of obtaining low atmo-
spheric CO2 stability levels will be significantly impacted by the combination of
biomass technology carbon capture and storage (Favero et al., 2020). It is techni-
cally possible to capture, transport, and repossess CO2 where biomass is utilized as
a fuel for gasification, combustion, and hydrogen generation in a big-scale plant.
Incorporating the charcoal produced by smaller, more scattered solid biomass gas-
ification operations into the soil has the potential to improve crop yields, soil water
retention, and soil carbon content (Bolan et al., 2022). People are suggesting these
alternatives to slow down climate change because they can quickly lower the amount
of CO2 in the air.

7 Biofuel Cost

It is now possible to produce a wide variety of biofuels. Conventional biofuels that


are considered “conventional” have seen extensive commercialization. These
include ethanol made from corn and sugar and fatty acid methyl ester biodiesel
made from oilseeds. Before biofuels with low net greenhouse gases and low indirect
land-use change impact are employed widely, a long way still needs to be traveled.
The crops that do not produce food and wastes obtained from lignocellulose (from
agriculture and forestry, including stoves, straws, bagasse, and woody biomass) are
used to create these fuels. Because electricity and hydrogen are not practical for use
in heavy-duty fleets or air travel, low-carbon biofuels are seen as essential to the
long-term decarbonization of the transportation industry. Carbon intensity limita-
tions in the United States and renewable fuel rules, like the federal Renewable Fuel
Standard (RFS), have promoted the adoption of both traditional and innovative bio-
fuels (LCFS). From 2011 to 2022, the RFS increased the requirements for renew-
able fuels to have a lower carbon footprint across their whole life cycle. The
definition of advanced fuels used in the RFS policy is different from the one used in
this article. It states that advanced fuels must cut greenhouse gas emissions over
their entire life cycle by at least 50% in comparison to a reference baseline based on
projections for 2010. The LCFS uses a continuous scale to measure the carbon
82 M. Irfan et al.

intensity of fuels and rewards those with a lower intensity without limiting the use
of alternative fuels like natural gas, electricity, or biofuels. The LCFS mandated a
gradual but steadily increasing reduction in carbon intensity in California’s trans-
portation fuel pool beginning in 2011 (Witcover & Williams, 2017).

8 Outlook

1. Unlike our staple foods, which have been refined over generations, committed
bioenergy crops have not been domesticated (Engels & Thormann, 2020). For
example, there are more than 600 species of eucalyptus, but it is not known how
to choose which is best for a particular location. If energy crops are to be grown
sustainably, it will be necessary to conduct additional research on the inputs of
agrochemicals and fertilizers.
2. Bioenergy crops need low-input systems that require only a small number of
resources such as water and fertilizer. Given the current state of breeding, signifi-
cant gains in bioenergy crop yields are projected to be achieved during the next
few decades. The cultivation of both genetically modified (GM) and non-GM
plants opens novel biotechnological avenues. Since the development of peren-
nial crops is less expensive and the soil’s chemistry and structure are preserved,
they are preferred over annual crops. Biomass yield is one metric, while plant
structure and chemical composition of the feedstock are others. C4 grasses, such
as miscanthus, may see increased productivity if their cold sensitivity is addressed
(McCalmont et al., 2017).
3. Developing new high-value energy sources from biomass could play a crucial
role in the industry’s long-term success. There are numerous biomaterials and
chemical feedstocks that may be derived from biomass, and they can also be
utilized as renewable hydrogen supply (Popa, 2018).
4. Going forward, biomass feedstocks will replace the petrochemicals used today.
Lubricants, textiles, polymers, biodegradable plastics, high matrix composites,
adhesives, stabilizers, paints, thickeners, and a variety of cellulosics are all
examples (Rahman et al., 2021). If high-value products can be recovered from
energy crops first, then the residues can be used to produce lower-value energy.
This could make energy crops more cost-effective. Food, animal feed, industrial
and chemical feedstocks, energy, and other items and materials could be made
from the many fractions of the total crop that are now being researched and
developed (de Jong & Jungmeier, 2015). For example, New Zealand built a pro-
totype plant with a closed loop to separate biomass into its constituent parts
(Krzeminski et al., 2017).
5  Promoting Energy Crops to Replace Fossil Fuel Use 83

9 Conclusions and Remarks

In this chapter, the potential of biofuel production from energy crops was described.
It is concluded that biofuel and biogas production from energy crops is an environ-
mentally friendly process. On the other hand, all kinds of bioenergy will immedi-
ately lower CO2 emissions when they are used in place of fossil fuels. Therefore, to
maximize the effects of carbon mitigation techniques, energy crop production
should be combined with carbon sinks and offset credits. However, it is believed
that one of the best methods to lessen the planet’s expanding ecological footprint is
through the usage of renewable energy. With specialized energy crops expected to
offer a higher share of biomass feedstock in the ensuing decades, bioenergy is
expected to continue its position as the largest contributor to global renewable
energy in the short to medium term and replace fossil fuel consumption. It is impor-
tant to conduct more research on the economics of producing biofuel because it is
helpful to control pollution, as it is an environmentally friendly approach.

References

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
lignocellulosic biomass. Bioresource Technology, 301, 122725. https://doi.org/10.1016/j.
biortech.2019.122725
Ahmad, N., Hamid, I., & Kazmi, S. T. H. (2017). Beyond COP21: What did Asian countries pledge
in the Paris Agreement.
André Cremonez, P., Feroldi, M., Cézar Nadaleti, W., de Rossi, E., Feiden, A., de Camargo,
M.  P., Cremonez, F.  E., & Klajn, F.  F. (2015). Biodiesel production in Brazil: Current sce-
nario and perspectives. Renewable and Sustainable Energy Reviews, 42, 415–428. https://doi.
org/10.1016/j.rser.2014.10.004
Aydın, H. (2016). Scrutinizing the combustion, performance and emissions of safflower biodiesel–
kerosene fueled diesel engine used as power source for a generator. Energy Conversion and
Management, 117, 400–409. https://doi.org/10.1016/j.enconman.2016.03.046
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting cir-cular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
84 M. Irfan et al.

Bolan, N., Hoang, S. A., Beiyuan, J., Gupta, S., Hou, D., Karakoti, A., Joseph, S., Jung, S., Kim,
K.-H., Kirkham, M. B., Kua, H. W., Kumar, M., Kwon, E. E., Ok, Y. S., Perera, V., Rinklebe,
J., Shaheen, S. M., Sarkar, B., Sarmah, A. K., et al. (2022). Multifunctional applications of
biochar beyond carbon storage. International Materials Reviews, 67(2), 150–200. https://doi.
org/10.1080/09506608.2021.1922047
Bórawski, P., Bełdycka-Bórawska, A., Szymańska, E.  J., Jankowski, K.  J., Dubis, B., &
Dunn, J.  W. (2019). Development of renewable energy sources market and biofuels in the
European Union. Journal of Cleaner Production, 228, 467–484. https://doi.org/10.1016/j.
jclepro.2019.04.242
Chandra, R., Takeuchi, H., & Hasegawa, T. (2012). Methane production from lignocellulosic agri-
cultural crop wastes: A review in context to second generation of biofuel production. Renewable
and Sustainable Energy Reviews, 16(3), 1462–1476. https://doi.org/10.1016/j.rser.2011.11.035
Cherry, C. E., & Pidgeon, N. F. (2018). Is sharing the solution? Exploring public acceptability of
the sharing economy. Journal of Cleaner Production, 195, 939–948. https://doi.org/10.1016/j.
jclepro.2018.05.278
Coady, M. D., Parry, I., Le, N.-P., & Shang, B. (2019). Global fossil fuel subsidies remain large:
An update based on country-level estimates. International Monetary Fund.
Czyżewski, B., Czyżewski, A., & Kryszak, Ł. (2019). The market treadmill against sustainable
income of European Farmers: How the CAP has struggled with Cochrane’s curse. Sustainability,
11(3), 791. https://doi.org/10.3390/su11030791
Datta, A., Hossain, A., & Roy, S. (2019). An overview on biofuels and their advantages and
disadvantages. Asian Journal of Chemistry, 31(8), 1851–1858. https://doi.org/10.14233/
ajchem.2019.22098
de Jong, E., & Jungmeier, G. (2015). Biorefinery concepts in comparison to petrochemical refiner-
ies. In Industrial biorefineries & white biotechnology (pp. 3–33). Elsevier.
Demirbaş, A. (2001). Biomass resource facilities and biomass conversion processing for fuels and
chemicals. Energy Conversion and Management, 42(11), 1357–1378. https://doi.org/10.1016/
S0196-­8904(00)00137-­0
Demirbas, A., Bafail, A., Ahmad, W., & Sheikh, M. (2016). Biodiesel production from non-­
edible plant oils. Energy Exploration and Exploitation, 34(2), 290–318. https://doi.
org/10.1177/0144598716630166
Desa, U. (2017). http://esa.un.org/unpp (gelesen am 16, 2010 (2008 revision). Accessed: May 19
(p. 2009b): World population prospects. United Nations Department of Economic and Social
Affairs Population Division.
DGS. (2005). Planning Installing bioenergy systems: A guide for installers. Architects and
Engineers, GREENPro project. Earthscan Publications.
Diallo, B., Li, M., Tang, C., Ameen, A., Zhang, W., & Xie, G. H. (2019). Biomass yield, chemical
composition and theoretical ethanol yield for different genotypes of energy sorghum culti-
vated on marginal land in China. Industrial Crops and Products, 137, 221–230. https://doi.
org/10.1016/j.indcrop.2019.05.030
Elavarasan, R.  M. (2019). The motivation for renewable energy and its comparison with other
energy sources: A review. European Journal of Sustainable Development Research, 3(1),
em0076. https://doi.org/10.20897/ejosdr/4005
Engels, J. M. M., & Thormann, I. (2020). Main challenges and actions needed to improve conser-
vation and sustainable use of our crop wild relatives. Plants, 9(8), 968. https://doi.org/10.3390/
plants9080968
Eryilmaz, T., Yesilyurt, M. K., Cesur, C., & Gokdogan, O. (2016). Biodiesel production potential
from oil seeds in Turkey. Renewable and Sustainable Energy Reviews, 58, 842–851. https://doi.
org/10.1016/j.rser.2015.12.172
Favero, A., Daigneault, A., & Sohngen, B. (2020). Forests: Carbon sequestration, biomass energy,
or both? Science Advances, 6(13), eaay6792. https://doi.org/10.1126/sciadv.aay6792
Fu, D.-h., Jiang, L.-y., Mason, A. S., Xiao, M.-l., Zhu, L.-r., Li, L.-z., Zhou, Q.-h., Shen, C.-j., &
Huang, C.-h. (2016). Research progress and strategies for multifunctional rapeseed: A case
5  Promoting Energy Crops to Replace Fossil Fuel Use 85

study of China. Journal of Integrative Agriculture, 15(8), 1673–1684. https://doi.org/10.1016/


S2095-­3119(16)61384-­9
Gabrielli, P., Gazzani, M., & Mazzotti, M. (2020). The role of carbon capture and utilization, car-
bon capture and storage, and biomass to enable a net-zero-CO2 emissions chemical industry.
Industrial and Engineering Chemistry Research, 59(15), 7033–7045. https://doi.org/10.1021/
acs.iecr.9b06579
Geissdoerfer, M., Vladimirova, D., & Evans, S. (2018). Sustainable business model innova-
tion: A review. Journal of Cleaner Production, 198, 401–416. https://doi.org/10.1016/j.
jclepro.2018.06.240
Grangeiro, L.  C., de Almeida, S.  G. C., de Mello, B.  S., Fuess, L.  T., Sarti, A., & Dussán,
K.  J. (2019). New trends in biogas production and utilization. In Sustainable bioenergy
(pp. 199–223). Elsevier.
Guo, S., Liu, Q., Sun, J., & Jin, H. (2018). A review on the utilization of hybrid renewable
energy. Renewable and Sustainable Energy Reviews, 91, 1121–1147. https://doi.org/10.1016/j.
rser.2018.04.105
Hajjari, M., Tabatabaei, M., Aghbashlo, M., & Ghanavati, H. (2017). A review on the prospects of
sustainable biodiesel production: A global scenario with an emphasis on waste-oil biodiesel uti-
lization. Renewable and Sustainable Energy Reviews, 72, 445–464. https://doi.org/10.1016/j.
rser.2017.01.034
Ilic, D., Williams, K., Farnish, R., Webb, E., & Liu, G. (2018). On the challenges facing the han-
dling of solid biomass feedstocks. Biofuels, Bioproducts and Biorefining, 12(2), 187–202.
https://doi.org/10.1002/bbb.1851
Islam, M. K., Wang, H., Rehman, S., Dong, C., Hsu, H. Y., Lin, C. S. K., & Leu, S. Y. (2020).
Sustainability metrics of pretreatment processes in a waste derived lignocellulosic biomass bio-
refinery. Bioresource Technology, 298, 122558. https://doi.org/10.1016/j.biortech.2019.122558
Kamm, B., Gruber, P. R., & Kamm, M. (2006). Biorefineries-industrial processes and products
(Vol. 2). Wiley-VCH Press.
Khatib, S. E., & Yassine, N. A. (2019). Advances in synthetic biology and metabolic engineer-
ing in the production of biofuel. International Journal of Current Microbiology and Applied
Sciences, 8(9), 1762–1772. https://doi.org/10.20546/ijcmas.2019.809.204
Krzeminski, P., Leverette, L., Malamis, S., & Katsou, E. (2017). Membrane bioreactors–a review
on recent developments in energy reduction, fouling control, novel configurations, LCA and
market prospects. Journal of Membrane Science, 527, 207–227. https://doi.org/10.1016/j.
memsci.2016.12.010
Li, W., & Zhang, H. (2017). Decomposition analysis of energy efficiency in China’s Beijing-­
Tianjin-­Hebei region. Polish Journal of Environmental Studies, 26(1), 189–203. https://doi.
org/10.15244/pjoes/65290
Liu, B., & Ji, S. (2013). Comparative study of fluidized-bed and fixed-bed reactor for syngas
methanation over Ni-W/TiO2-SiO2 catalyst. Journal of Energy Chemistry, 22(5), 740–746.
https://doi.org/10.1016/S2095-­4956(13)60098-­4
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In S. A. Bandh (Ed.),
Climate change: The social and scientific construct. Springer Nature Publishing. https://link.
springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
McCalmont, J. P., Hastings, A., McNamara, N. P., Richter, G. M., Robson, P., Donnison, I. S., &
Clifton-Brown, J. (2017). Environmental costs and benefits of growing Miscanthus for bio-
energy in the UK. Global Change Biology Bioenergy, 9(3), 489–507. https://doi.org/10.1111/
gcbb.12294
Molino, A., Chianese, S., & Musmarra, D. (2016). Biomass gasification technology: The state
of the art overview. Journal of Energy Chemistry, 25(1), 10–25. https://doi.org/10.1016/j.
jechem.2015.11.005
Muntean, M., Guizzardi, D., Schaaf, E., Crippa, M., Solazzo, E., Olivier, J., & Vignati, E. (2018).
Fossil CO2 emissions of all world countries (p. 2). Publications Office of the European Union.
86 M. Irfan et al.

Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Nelson, L., Park, S., & Hubbe, M. A. (2018). Thermal depolymerization of biomass with emphasis
on gasifier design and best method for catalytic hot gas conditioning. BioResources, 13(2),
4630–4727. https://doi.org/10.15376/biores.13.2.Nelson
Nikkhah, A., El Haj Assad, M., Rosentrater, K.  A., Ghnimi, S., & Van Haute, S. (2020).
Comparative review of three approaches to biofuel production from energy crops as feed-
stock in a developing country. Bioresource Technology Reports, 10, 100412. https://doi.
org/10.1016/j.biteb.2020.100412
Owusu, P. A., & Asumadu-Sarkodie, S. (2016). A review of renewable energy sources, sustain-
ability issues and climate change mitigation. Cogent Engineering, 3(1), 1167990. https://doi.
org/10.1080/23311916.2016.1167990
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Petit-Boix, A., & Leipold, S. (2018). Circular economy in cities: Reviewing how environmental
research aligns with local practices. Journal of Cleaner Production, 195, 1270–1281. https://
doi.org/10.1016/j.jclepro.2018.05.281
Popa, V. I. (2018). Biomass for fuels and biomaterials. In Biomass as renewable raw material to
obtain bioproducts of high-tech value (pp. 1–37). Elsevier.
Popp, J., Harangi-Rákos, M., Gabnai, Z., Balogh, P., Antal, G., & Bai, A. (2016). Biofuels and their
co-products as livestock feed: Global economic and environmental implications. Molecules,
21(3), 285. https://doi.org/10.3390/molecules21030285
Prasad, S., Kumar, S., Sheetal, K., & Venkatramanan, V. (2020). Global climate change and biofuels
policy: Indian perspectives. In Global climate change and environmental policy (pp. 207–226).
Springer.
Qian, K., Kumar, A., Patil, K., Bellmer, D., Wang, D., Yuan, W., & Huhnke, R. L. (2013). Effects
of biomass feedstocks and gasification conditions on the physiochemical properties of char.
Energies, 6(8), 3972–3986. https://doi.org/10.3390/en6083972
Rahman, M. S., Hasan, M. S., Nitai, A. S., Nam, S., Karmakar, A. K., Ahsan, M. S., Shiddiky,
M. J. A., & Ahmed, M. B. (2021). Recent developments of carboxymethyl cellulose. Polymers,
13(8), 1345. https://doi.org/10.3390/polym13081345
Raman, L. A., Deepanraj, B., Rajakumar, S., & Sivasubramanian, V. (2019). Experimental inves-
tigation on performance, combustion and emission analysis of a direct injection diesel engine
fuelled with rapeseed oil biodiesel. Fuel, 246, 69–74. https://doi.org/10.1016/j.fuel.2019.02.106
Rashid, A., Asif, F. M. A., Krajnik, P., & Nicolescu, C. M. (2013). Resource conservative manufac-
turing: An essential change in business and technology paradigm for sustainable manufactur-
ing. Journal of Cleaner Production, 57, 166–177. https://doi.org/10.1016/j.jclepro.2013.06.012
Roy, S., Dikshit, P. K., Sherpa, K. C., Singh, A., Jacob, S., & Chandra Rajak, R. C. (2021). Recent
nanobiotechnological advancements in lignocellulosic biomass valorization: A review. Journal
of Environmental Management, 297, 113422. https://doi.org/10.1016/j.jenvman.2021.113422
Rozina, Asif, S., Ahmad, M., Zafar, M., & Ali, N. (2017). Prospects and potential of fatty acid
methyl esters of some non-edible seed oils for use as biodiesel in Pakistan. Renewable and
Sustainable Energy Reviews, 74, 687–702. https://doi.org/10.1016/j.rser.2017.02.036
Rutz, D., & Janssen, R. (2007). Biofuel technology handbook (p. 95). WIP Renewable Energies.
Scarlat, N., Dallemand, J.-F., & Fahl, F. (2018). Biogas: Developments and perspectives in Europe.
Renewable Energy, 129, 457–472. https://doi.org/10.1016/j.renene.2018.03.006
Schipfer, F. (2017). Densification and conversion technologies for bioenergy and advanced bio-
based material supply chains-a European case study. TU Wien.
Searchinger, T. D., Beringer, T., Holtsmark, B., Kammen, D. M., Lambin, E. F., Lucht, W., Raven,
P., & van Ypersele, J.-P. (2018). Europe’s renewable energy directive poised to harm global
forests. Nature Communications 9, 3741. https://doi.org/10.1038/s41467-­018-­06175-­4
5  Promoting Energy Crops to Replace Fossil Fuel Use 87

Sirajunnisa, A.  R., & Surendhiran, D. (2016). Algae–A quintessential and positive resource of
bioethanol production: A comprehensive review. Renewable and Sustainable Energy Reviews,
66, 248–267. https://doi.org/10.1016/j.rser.2016.07.024
Suresh, M., Jawahar, C.  P., & Richard, A. (2018). A review on biodiesel production, combus-
tion, performance, and emission characteristics of non-edible oils in variable compression ratio
diesel engine using biodiesel and its blends. Renewable and Sustainable Energy Reviews, 92,
38–49. https://doi.org/10.1016/j.rser.2018.04.048
Thangavelu, S. K., Ahmed, A. S., & Ani, F. N. (2016). Review on bioethanol as alternative fuel for
spark ignition engines. Renewable and Sustainable Energy Reviews, 56, 820–835. https://doi.
org/10.1016/j.rser.2015.11.089
Usman, M., Makhdum, M. S. A., & Kousar, R. (2021). Does financial inclusion, renewable and
non-renewable energy utilization accelerate ecological footprints and economic growth? Fresh
evidence from 15 highest emitting countries. Sustainable Cities and Society, 65, 102590.
https://doi.org/10.1016/j.scs.2020.102590
Usmani, Z., Sharma, M., Karpichev, Y., Pandey, A., Chander Kuhad, R. C., Bhat, R., Punia, R.,
Aghbashlo, M., Tabatabaei, M., & Gupta, V.  K. (2020). Advancement in valorization tech-
nologies to improve utilization of bio-based waste in bioeconomy context. Renewable and
Sustainable Energy Reviews, 131, 109965. https://doi.org/10.1016/j.rser.2020.109965
Van Duren, I., Voinov, A., Arodudu, O., & Firrisa, M. T. (2015). Where to produce rapeseed bio-
diesel and why? Mapping European rapeseed energy efficiency. Renewable Energy, 74, 49–59.
https://doi.org/10.1016/j.renene.2014.07.016
Villadsen, S. N. B., Fosbøl, P. L., Angelidaki, I., Woodley, J. M., Nielsen, L. P., & Møller, P. (2019).
The potential of biogas; the solution to energy storage. ChemSusChem, 12(10), 2147–2153.
https://doi.org/10.1002/cssc.201900100
Wakil, M. A., Kalam, M. A., Masjuki, H. H., Atabani, A. E., & Rizwanul Fattah, I. M. (2015).
Influence of biodiesel blending on physicochemical properties and importance of mathematical
model for predicting the properties of biodiesel blend. Energy Conversion and Management,
94, 51–67. https://doi.org/10.1016/j.enconman.2015.01.043
Witcover, J., & Williams, R. B. (2017). Biofuel tracker: Capacity for low carbon fuel policies–
assessment through 2018.
Wu, Y., Yang, W., & Blasiak, W. (2014). Energy and exergy analysis of high temperature agent
gasification of biomass. Energies, 7(4), 2107–2122. https://doi.org/10.3390/en7042107
Zhang, Z.-H., & Balasubramanian, R. (2016). Investigation of particulate emission characteristics
of a diesel engine fueled with higher alcohols/biodiesel blends. Applied Energy, 163, 71–80.
https://doi.org/10.1016/j.apenergy.2015.10.173
Zhao, Y. L., Dolat, A., Steinberger, Y., Wang, X., Osman, A., & Xie, G. H. (2009). Biomass yield
and changes in chemical composition of sweet sorghum cultivars grown for biofuel. Field
Crops Research, 111(1–2), 55–64. https://doi.org/10.1016/j.fcr.2008.10.006
Zieliński, M., Dębowski, M., Kisielewska, M., Nowicka, A., Rokicka, M., & Szwarc, K. (2019).
Cavitation-based pretreatment strategies to enhance biogas production in a small-scale agricul-
tural biogas plant. Energy for Sustainable Development, 49, 21–26. https://doi.org/10.1016/j.
esd.2018.12.007
Chapter 6
Changes in the Agriculture Sector That
Are Essential to Mitigate and Adapt
to Climate Changes

Enohetta B. Tambe, Charles C. Anukwonke, Iheoma E. Mbuka-Nwosu,


and Chinedu I. Abazu

Abstract  Growing human population and preference for diet are driving global
food demands. Consequently, the environmental systems that sustain food produc-
tion is being stretched, altering agricultural networks and leading to an increasing
number of people living in hunger in some regions across years. The situation is
exacerbated by climate change, which is mainly caused by greenhouse gas emis-
sions through anthropogenic activities, including the agricultural sector. Sustaining
the environmental systems for food production requires climate-smart contributions
in the agricultural sector, which should principally achieve increase in food produc-
tivity, enhance resilience and reduce emissions. These contributions are essential to
address production of better, nutritious and more food and pest and disease manage-
ment, policies and programmes that reduce farmers’ vulnerability to shocks, mak-
ing data available for informed decision-making, reduce deforestation, strengthen
carbon sequestration, reduce emission per kilogram of food produced and insure
farmers among others. However, accessing and accommodating these climate-smart
agricultural (CSA) changes is not uniform across regions, especially in low-income
economies that are mainly dependent on climate-sensitive livelihoods. These limita-
tions challenge the progress and well-being of these people who are already suffer-
ing existing deprivations. While the global temperature remains dynamic, managing
the interconnectedness, necessity and traditional techniques of these CSA changes
with the quest for progress in food security is an essential consideration. Therefore,
a nexus paradigm that strengthens synergies and accommodates the trade-offs of

E. B. Tambe · C. C. Anukwonke (*)


Department of Environmental Management, Chukwuemeka Odumegwu Ojukwu University,
Uli, Nigeria
I. E. Mbuka-Nwosu
Department of Environmental Management, Federal University of Technology,
Owerri, Nigeria
C. I. Abazu
Department of Urban and Regional Planning, Chukwuemeka Odumegwu Ojukwu University,
Uli, Nigeria

© The Author(s), under exclusive license to Springer Nature 89


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_6
90 E. B. Tambe et al.

these CSA changes is a viable option for mitigation and adaptation, while taking
into cognizance local priorities in realising sustained food needs.

Keywords  Agricultural changes · Climate-smart agriculture · Synergies ·


Mitigation · Adaptation

1 Introduction

Ending hunger in the growing human population is one of the contemporary chal-
lenges facing societies. In the course of meeting food needs for the growing popula-
tion, the networks on which survival depends are being weakened. This is occurring
through agricultural production at the expense of huge transformation of environ-
mental resources, such as water, soils, biodiversity and natural resources (Brack,
2019; Klauser, 2021). Following this recurrent transformation of environmental
resources and subsequent creation of conditions such as spread of transboundary
diseases and pests, production of copious volumes of wastes and their regenerative
capacities are stretched, leading to unhealthy environmental support systems (FAO,
2017a; Tambe et al., 2022). The consequence of this is a vicious cycle that under-
mines productivity and reduces agricultural output and its interconnected impacts
that lead to hunger, poverty, deprivation and human suffering (Anukwonke et al.,
2022). Over the years, climate change has posed as a true titian with one of such
challenge emanating from the agricultural sector is the emission of greenhouse
gases such as methane, nitrous oxide and carbon dioxide into the atmosphere (Lynch
et al., 2021). These emissions are estimated to contribute between 19% and 29% of
total anthropogenic greenhouse gas emissions (CCAFS, 2015; Klauser, 2021).
The addition of these gases causes enhanced increase of global temperature, a
condition usually referred to as ‘climate change’ (IPCC, 2021). This environmental
condition leads to alteration of ecological processes, shifting of agroecosystem
boundaries, weather variability, invasive species and frequent extreme weather
events. Over the years, climate shifts have manifested as a true titian with concerns
on the major source of livelihood and stresses on soil fertility (Akanwa et al., 2019;
Anabaraonye et al., 2021). These environmental stressors weaken human survival
networks and undermine achievement of all the sustainable development goals
(Anukwonke et al., 2022; United Nations, 2017; Malla et al., 2022; Bandh et al.,
2021, 2023; Mushtaq et al., 2020). As a result of the deleterious effects of climate
change and the contribution from the agricultural sector, it is essential to identify the
pathways through which the sector is contributing to greenhouse gas emissions.
This identification facilitates restructuring of these pathways with a view of achiev-
ing networks that could curb these emissions and sustain human progress (Akinnagbe
& Irohibe, 2014; Mumtaz et al., 2019; Parray et al., 2022 Bandh et al., 2022 Bandh,
2022a, b).
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 91

The necessity to restructure these pathways is based on the premise that through-
out human agricultural history, experts have recently come to understand that the
focus on agriculture, which has been on investment in resource-intensive farming
systems and high input, cannot guarantee sustainable agriculture and food produc-
tion for society (FAO, 2017a). Consequently, ‘holistic’ approaches are needed that
take into cognizance climate-smart agriculture (CSA), agroforestry, conservation
agriculture and agro-ecology. Each of these approaches seek to address sustainable
resource use and attainment of at least one of these core pillars: increased food pro-
ductivity, enhanced resilience and reduced emissions. Following the potential of
CSA to address all of these pillars, many authors contend that climate-smart agri-
culture incorporates many of the field-based and farm-based sustainable agricultural
land management strategies that are already well known and in widespread use,
including residue management, agroforestry and conservation tillage (Abhilash
et al., 2021; FAO, 2010). Consequently, this approach (CSA), which takes into con-
sideration traditional and indigenous knowledge, has become the paradigm for sus-
tainable agriculture in the twenty-first century and is detailed in this chapter.
However, while these approaches are promising tools in sustaining the agricul-
tural sector, the pathways and resources to achieve their pillars vary and are not
uniformly accessible across regions (Anuga et al., 2019; CIAT & World Bank, 2018;
Opeyemi et al., 2021). The variation necessitates a consideration of local priorities
and nexus approaches that capture significant synergies and accommodate trade-­
offs (Abegunde & Obi, 2022). Similarly, lagging of its utilisation and acceptability
is characterised in economies that are worst hit by food insecurity and with huge
reliance on climate-sensitive livelihoods. This challenges the existing poor eco-
nomic status of these people and undermines achievement of sustainable develop-
ment goals. Exploring the frontiers of the extent of application of these approaches,
their efficiency, acceptability and benefits could provide opportunity to widen their
scope and promote a sustainable pattern of agriculture. However, while the climate
is likely to keep changing, and more food is needed to feed the ever-growing human
population, some of the approaches are likely to remain dynamic to accommodate
these changes. In this regard, to what extend can we keep changing these approaches
while sustaining the pillars they seek? These are pertinent issues that challenge the
future we want. This chapter has addressed the explored changes in the agricultural
sector necessary to sustain food production in the midst of accelerating climate and
examine issues that open frontiers for investigation in sustaining the sector.
92 E. B. Tambe et al.

2 Feeding the World in Climate Crisis: The Sources


of Agricultural Greenhouse Gases

2.1 Food Demands and Productivity

Projection on feeding and meeting increasing caloric needs of the growing human
population is a shared vision for member countries of the Food and Agriculture
Organization (FAO). Despite progress made so far to reduce food insecurity, the
challenges of meeting global food needs are rife. There is illuminating inequality,
about 795 million people still suffer from hunger and more than two billion are fac-
ing micronutrient deficiencies (FAO, 2017a; OPHI & UNDP, 2019). Consequently,
achieving food security in the next decades is quite challenging in the face of further
deforestation, land degradation and climate change (FAO, 2016). Notwithstanding
the increasing global climate, food production with reference to the first decade of
the twenty-first century is expected to increase by 70% in 2050 to cater to the pro-
jected 9–10 billion people in the world. In developing countries, their production is
expected to double (FAO, 2009, 2017a). Similarly, 90% (with over 80% in develop-
ing countries) of the expected increase in agricultural production will employ
increased intensive agriculture, while 10% will make use of land expansion. While
crop yield is expected to grow, the growth rate will be at a slower rate (decelerating
growth) when compared to historical records.
The overall agricultural production is expected to strengthen livelihood assets,
enhance the well-being status of farmers and the rural poor and reduce the preva-
lence of undernourishment in developing countries from 16.3% in the first decade
of the twenty-first century to 4.8% by 2050 (FAO, 2009). This is aligned to the fact
that in many economies across the world, agriculture makes up a sizeable portion of
the GDP, and 2.5 billion people globally rely on it for their living (FAO, 2016).
Similarly, given that roughly two-thirds of the world’s poorest people work in agri-
culture and that three-quarters of the world’s poorest people still reside in rural
regions, improving global agricultural performance is essential to alleviating pov-
erty and food insecurity. These achievements in relation to the role of agriculture
seem impressive and are interconnected to the pathways of realising a sustain-
able future.
Unfortunately, the projected increase in food production with its associated
expansion of land use by 10% will occur at the expense of reducing the forest can-
opy (Brack, 2019). Although there is decelerating deforestation associated with
agricultural development, agriculture is the main driver of forest removal and
accounted for 27% of all tree canopy loss between 2001 and 2015 (Brack, 2019;
European Commission, 2013). However, there are variations in contributions of this
forest canopy loss across the various agricultural sectors. According to the European
Commission (2013), livestock accounted for 46%, crops for animal feed 11%, soy-
beans 19%, oil palm 8%, maize 11%, sugarcane 5% and rice 6%. The replacement
of forest areas with these practices in an unsustainable way increases the vulnerabil-
ity of the practices to environmental stressors. Increased temperatures, weather
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 93

unpredictability, shifting agro-ecosystem boundaries, invasive plants and pests and


an increase in extreme weather events are just a few of the negative effects of cli-
mate change that are already being felt (World Bank, 2021b). On farms, climate
change is diminishing animal output, the nutritional value of main grains and crop
yields. To sustain existing yields and enhance output and food quality to satisfy
demand, significant adaptation and mitigation expenditures will be needed
(UNFCCC, 2021; World Bank, 2021a).

2.2 Agriculture and Greenhouse Gas Emissions: Land


Preparation to Pre-harvest

Expanding land use in agriculture is reducing forest canopy and degrading its envi-
ronment. This reduction in forest canopy emanating from the agricultural sector is
posing huge challenges in sustaining agricultural production and a healthy environ-
ment for societies (Brack, 2019; FAO, 2016). This is based on the premise that the
forests are responsible for capturing a significant quantity of carbon dioxide from
the atmosphere. The captured carbon dioxide is used during photosynthesis (food
production in the plant) and constitute the biomass of the plant: roots, leaves,
branches and tree trunks. Carbon is also stored in forests in soils, roots, leaf litter
and woody debris. These carbon-storing sectors of the forests and their environment
are essential in mitigating climate change from anthropogenic activities.
When the forests and their environment are replaced through agricultural devel-
opment that cannot balance the carbon capture, the result is decrease in the sources
of carbon sinks to sustain a healthy environment, increase greenhouse gases in the
atmosphere, increase erosion and siltation, loss of freshwater resources and a weak-
ening of a myriad of ecological networks that sustain life (Ali et al., 2020; Ancha
et al., 2019). These in turn undermine the support system of agricultural production
and increase the climate challenge – a vicious cycle, characterised by poverty, hun-
ger, migration, inequality, gender inequality, urban challenges and violent conflict
among others that undermine progress in societies (Anukwonke et al., 2022).
While the agricultural sector is an essential tenet in sustaining human dignity, the
practice is also contributing hugely (estimated to be 19–29%) to the net total anthro-
pogenic greenhouse gas emissions such as methane, carbon dioxide and nitrous
oxide (Crippa et al., 2021; Lynch et al., 2021; UNFCCC, 2021). The main sources
of greenhouse gas emissions in the sector usually occur during agricultural opera-
tion and inputs, such as the use of tractor fuel, fertilizer manufacture and applica-
tion, soil tillage, transportation, production and transportation of pesticide, threshing,
harvesting and irrigation; agricultural land management such as manure manage-
ment, waste management and N fertilizer management; and biomass burning and
decay among others (Kitamura et  al., 2021; Lazcano et  al., 2021; Tuğrul, 2019).
Greenhouse gas emission in the sector also occurs through food losses in the pro-
duction chain from the harvest to table (FAO, 2014, 2015). For example, when
94 E. B. Tambe et al.

chemical fertilizers are applied on soils to enhance plant growth, a series of chemi-
cal reactions lead to the emission of nitrous oxide, contributing about 2.5% of total
greenhouse gas emissions (Lazcano et al., 2021). Consequently, expanding agricul-
tural lands through clearing the forest canopy (and replacing with crops and/or live-
stock) while seeking to maintain a habitable greenhouse gas emission requires
identifying, mimicking and integrating natural approaches in the agricultural sector
that should maintain the values of the natural land use (natural forests) or enhance it.

2.3 Food Losses and Greenhouse Gas Emissions: Harvest


to Consumers’ Table

It is estimated that 33% of all foods produced across the world for human consump-
tion do not reach the consumer’s table (FAO, 2014, 2015). It ends up as food wastes.
While this remains a challenge to the economy, food security and environmental
health, it is also a waste of natural resources (Łaba et al., 2022; World Bank, 2020).
This is because, in the course of food production, energy and resources are being
invested for growing, processing, packaging, transportation and marketing the food.
The utilisation of energy by tractors to prepare the land for cultivation, energy used
by distribution vehicles and industrial processes emit a copious amount of carbon
dioxide into the environment. Similarly, in the production and usage of input mate-
rials such as fertilizers, irrigation process and packaging materials, emission of
greenhouse gases occurs. Furthermore, the degradation of food waste is a potent
source of methane emission. Globally, these emissions associated with food waste
is estimated at 8% (4.4GtCO2eq) of the total anthropogenic greenhouse gas emis-
sions (FAO, 2015; World Bank, 2020).
While food losses occur at all stages from harvest to consumers’ table, the vol-
ume of losses varies in space and depends on the local condition of each economy
and income status (FAO, 2015; Obinaju & Ikpeida, 2021; Seberini, 2020). The per
capita footprint of food losses is highest in North America and is least in sub-­
Saharan Africa. Generally, high-income economies suffer more food losses at the
processing, transportation and consumption stages. This has been attributed to aes-
thetic preferences. For low-income economies, higher food losses occur at the pro-
duction and post-harvest phases (Obinaju & Ikpeida, 2021; World Bank, 2020).
This has been correlated with weak infrastructure and insufficient knowledge on
storage and handling in the climatic conditions that favours food spoilage.
The losses could vary for similar crops and livestock across regions (Barthelmie,
2022; FAO, 2015; Mrowczyńska-Kamińska et  al., 2021). For example, in the
European economies, vegetable production is more carbon-intensive (grams of car-
bon dioxide associated with the production of one unit of electricity in Kw/h) than
its related production in the industrialised and Southeast Asia. Conversely, Asia is
more carbon intensive in cereal production than Europe. The difference is associ-
ated with the type of cereal produced. For example, because of the decomposition
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 95

of organic matter in paddy fields, rice (commonly grown in Asia) has a higher car-
bon intensity than wheat.
Furthermore, while the volume of loss of some foods like meat could be rela-
tively small, their contributions to climate impact, expressed in carbon footprint, are
huge when compared to plant-based food (FAO, 2015; Poore & Nemecek, 2018).
The carbon footprint of food waste is the total amount of greenhouse gases emitted
throughout its life cycle, and it is measured in kg of CO2 equivalents. For example,
meat contributes less than 5% of total food losses. Unfortunately, it contributes
more than 20% of the carbon footprint associated with food losses. This is because
the calculation of the carbon footprint of meat waste takes into consideration all the
processes that lead to meat production. These include fertilizer used to produce the
feed, emission from ruminants and those related to manure management.
Consequently, it is a cause for concern and necessitates that emission reduction
strategies should focus on major climate hotspot foods such as cereals and meat.
Also, although the consumption phase of the food production chain accounts
only about 22% of total food losses, it contributes to 37% of the carbon footprint
associated with food losses (FAO, 2015). This is because the closer the food is to the
table (e.g. 1 kg of rice), any loss will have a higher carbon intensity associated with
harvesting, transportation and processing compared to a similar quantity of food
lost at an earlier stage (e.g. 1 kg of rice during harvesting) of the production chain.
Tracing these pathways that greenhouse gas emissions occur in the food production
chain constitutes an essential prerequisite in restructuring the agricultural sector
with a view of imbibing ways of addressing the weaknesses of the supply chain,
curbing emissions and achieving a productive agricultural system.

3 Agricultural Changes Necessary to Mitigate and Adapt


to Climate Change

3.1 Concepts and Principles for Sustainable Agriculture:


Climate-Smart Agriculture

In the course of human quest in achieving sustainability in agriculture and a produc-


tive society, it has become clearer that ‘holistic’ approaches are needed that take
into cognizance climate-smart agriculture (CSA), agroforestry, conservation agri-
culture and agro-ecology (Klauser, 2021; Leakey, 2019). Many authors contend that
climate-smart agriculture incorporates many of the field-based and farm-based sus-
tainable agricultural land management strategies that are already well known and in
widespread use, including residue management, agroforestry and conservation till-
age (Abhilash et  al., 2021; FAO, 2010). This is because the application of these
fields and farm practices, as well as how they might be enhanced in light of a chang-
ing climate, has, meanwhile, received the majority of attention in climate-smart
agriculture.
96 E. B. Tambe et al.

Climate-smart agriculture (CSA), according to FAO, is a strategy for changing


food and agricultural systems to promote sustainable development and protect food
security in the face of climate change (FAO, 2019a). It is a crucial method of
addressing the interconnected problems of food security and climate change by
managing the landscapes’ crops, livestock, forests and fisheries (World Bank,
2018a, 2021a). To maintain crop output, CSA can contribute to adaptation and miti-
gation techniques. For instance, CSA improves soil properties, water and fertilizer
usage efficiency, and yield stability and strengthen carbon capture, all of which help
to produce agricultural systems that are more climate resilient (Abhilash et  al.,
2021; FAO, 2019b).
CSA was introduced at the 2010 Hague Conference on Agriculture, Food
Security, and Climate Change and is a pathway to contribute to the attainment of
sustainable progress (Amin et al., 2015; Klauser, 2021). This contemporary pattern
of agriculture is based on the premise of achieving triple success: increased food
productivity, enhanced resilience and reduced emissions. Although CSA is based on
existing technology, knowledge and principles of sustainable agriculture, it is dis-
tinct in some aspects. The approach captures the pathways in the agricultural pro-
duction chain (cradle to grave) that address emissions, such as food losses; considers
the synergies and trade-offs between productivity, mitigation and adaptation; and
takes into cognizance new funding opportunities and reduces the deficiency in
investment. It addresses the interlinked challenge of accelerating climate change
and food security (availability, stability, access and utilisation).
While ensuring environmental health, CSA seeks to sustain the quality of food
produced and maintain an acceptable quantity of food and cultural satisfaction of
societies. This in turn is contributing in reducing vulnerability to poverty in societ-
ies and hunger and enhancement of fullness of life, especially for those with huge
reliance on climate-sensitive livelihoods. In this regard, CSA falls within the con-
fines of sustainability and captures most of the sustainable development goals and
their interconnectedness, such as goal number 2 on zero hunger, 4 on descent job
and economic growth, 13 on climate action, 12 on responsible consumption and
production and 14/15 on life on land and below water (Anukwonke et  al., 2022;
Klauser, 2021). In achieving the triple success of CSA, trade-offs must frequently
be made (FAO, 2017a). In order to do this, we must consider the pathways that
hamper the agricultural system, find synergies and evaluate the advantages and dis-
advantages of various choices in light of the stakeholder objectives discovered
through participatory methods.
In addition, CSA upholds ecosystem services and adopts a landscape strategy
that expands on the ideas of sustainable agriculture while moving beyond the con-
strained sectoral strategies that lead to competing and uncoordinated land uses and
toward integrated planning and management (CCAFS, 2022; FAO, 2017a). While
CSA is a method for directing measures to change agri-food systems toward envi-
ronmentally friendly and climate-resilient practices, it also aids in achieving other
globally recognised objectives like the Sustainable Development Goals (SDGs) and
the Paris Agreement (UNFCCC, 2021). CSA supports the FAO Strategic Framework
2022–2031, which is anchored on the ‘four betters’  – better productivity, better
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 97

nutrition, better environment and a better living for everyone, leaving no one behind.
The strategic framework of FAO suggests that the strategy be implemented through
five action points: boosting the funding and finance alternatives, strengthening
national and local institutions, supporting enabling policy frameworks and applying
CSA practices at the field level.
Many of the strategies that make up CSA already exist globally and are utilised
by farmers to manage a variety of production risks, despite the fact that the idea is
new and continually evolving (CCAFS, 2022). In order to mainstream CSA, a thor-
ough inventory of current practices, future-looking behaviours and institutional and
financial enablers for CSA adoption must be done. The creation of technologies and
practices, the production of climate change models and scenarios, information tech-
nologies, insurance plans, value chains and the improvement of institutional and
political enabling environments are just a few of the many entrance points for
CSA. It therefore incorporates various interventions at the food system, landscape,
value chain or policy level in addition to just a few technologies at the farm level.
In addition to providing potential for agricultural sector development and eco-
nomic progress, CSA technology and practices offer ways to address the difficulties
posed by climate change. Consequently, a practice is said to be CSA if it advances
food security and at least one of the other CSA goals of mitigation and/or adaptation
to climate change (FAO, 2018a). Around the world, CSA refers to thousands of
technologies and methods. According to evidence from the literature, farmers are
utilising a number of agricultural innovations either from native knowledge or new
technology, such as conservation agriculture, intercropping/crop diversification, ter-
racing, enhanced seedling and integrated soil fertility management to increase their
capacity for climate change and variability adaptation. The variation of these meth-
ods and their application means a CSA method should be created in a context-­
specific way, taking into consideration local climatic and environmental, market,
economic and cultural variables, in order to be as successful as feasible
(Celeridad, 2018).

3.2 The Action Plans for Realising CSA: Mitigation


and Adaptation Strategies

Around the world, there is growing interest in CSA interventions. These interven-
tions take into consideration different elements such as policies, investments, insti-
tutions and technologies, harmonised in local contexts and applicable off-farm and
on-farm (FAO, 2021; Matteoli et al., 2021). These interventions are usually referred
to as ‘mitigation and adaptation and strategies’. While mitigation proffers ways to
reduce the impacts of climate change, adaptation seeks to chart pathways to accom-
modate and live with the changing climate. These strategies are mutually applied
and have been evolving (Rasul & Sharma, 2016). To facilitate this in agriculture, the
FAO’s five action plans for CSA implementation, which mitigate and adapt to
98 E. B. Tambe et al.

climate change and ensure sustainable food production, serve as the foundation to
achieve the triple win of CSA: increased productivity, enhanced resilience and
reduced emissions. These action plans are as follows: widen the base of evidence,
support planning and enabling policy frameworks, boost local and national institu-
tions, expand available financing and put procedures into action in the field.
Although the various agricultural sub-sectors face specific challenges in achieving
the triple win of CSA, these sub-sectors are interconnected (Matteoli et al., 2021).
Thus, understanding this interconnectedness provides better tools to strengthen syn-
ergies within and across the sub-sectors, while reducing the trade-offs. These con-
siderations constitute the highlights of this section.
Firstly, a sound evidence base is a crucial enabling factor in developing agricul-
tural policy. This should capture the current and projected impact of climate change
in the specific economy and identify key areas of vulnerabilities (FAO, 2017a;
Rosenstock et al., 2019). Therefore, a key component of enhancing nations’ ability
for adaptation is expanding the evidence base. Assessing the effects of climate
change and greenhouse gas (GHG) emissions from agriculture and food systems,
finding and analysing climate-smart alternatives within the context of sustainable
development, determining the institutional and financial requirements for imple-
mentation and, of course, information obtained through the feedback loop of moni-
toring and evaluation are just a few examples.
Governments should be able to respond to inquiries like these with the help of
the information gathered. This information gathered should respond to questions
such as the following: What are the most probable climatic consequences at the
sectoral and sub-sectoral levels, and how do the time frames connected to those
impacts influence the timeline of adaptation interventions? What is the relationship
between the costs of adaptation for a particular sector or sub-sector and the revenues
and benefits to livelihoods that are anticipated to result? Would boosting imports
and diversifying the local, regional and national economy be a better use of these
investments? The promotion of a gender-responsive strategy is one significant part
of developing the evidence basis, which is a fundamental emphasis of climate-smart
agriculture (FAO, 2021). As a result, this stage should involve gathering data that
has been broken down by sex and doing a gender analysis. It serves as the founda-
tion for agricultural policies and project design that encourage the equality and
equity of opportunity for men and women in this way (FAO, 2017a).
After a strong evidence base is established, a unified national CSA plan (policy)
is created. This plan serves as a guide for integrating agriculture-related climate
change measures into relevant sectoral plans and strategies; existing policies are
revised and new policies are developed as necessary to create the right conditions
for implementing prioritised CSA options, set the right incentives, remove obstacles
to adoption and account for potential trade-offs (FAO, 2017a). It is necessary to
perform a thorough analysis of present policies and their intended and unforeseen
effects on the top goals for national development. The objective is to provide guid-
ance for the modification and development of policies and guarantee the highest
level of policy coherence, supported by inclusive decision-making procedures and
multi-stakeholder discussion.
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 99

Coordinated planning and policy coherence across sectors allow for the identifi-
cation of pertinent interlinkages, the enhancement of synergies and the avoidance,
at the very least minimisation, or compensation of potential trade-offs (Lewis &
Rudnick, 2019). There may be trade-offs between current policies enacted to
increase the priority area of agricultural production and goals for adaptation and
mitigation as well as for the sustainable use of natural resources. For instance, pro-
viding fossil fuel or energy subsidies to pump irrigation water in arid areas may
temporarily boost output. However, in the long run, they may lead to energy loss
and overuse of water resources, which would ultimately weaken farmers’ resilience
(FAO, 2017a). Similarly, increasing the share of bioenergy usage as an alternative to
fossil in agricultural systems should consider the huge water needs associated with
this renewable energy option (Liu et  al., 2018). Thus, synergies and trade-offs
should take into consideration several parameters, especially local priorities.
The tenure rights of food producers are a crucial factor to take into account when
planning climate-smart agriculture operations (FAO, 2017a). Numerous techniques
endorsed by CSA programmes, such agroforestry or conservation agriculture, need
upfront investment and take time to reap the rewards. The only way to ensure that
agricultural producers would profit from such investments is to have solid tenure
rights to cultivated land. Adopting advocated methods is sometimes hampered by
weak or absent tenure rights since there is a danger of being evicted, especially for
women and indigenous people.
Social protection programmes and initiatives may be significant components of
the CSA implementation process and contribute to national efforts on social protec-
tion and equality, depending on the socio-economic background and preferred
climate-­smart agricultural alternatives (FAO, 2017a; FAO, 2018b). Food vouchers,
cash transfers, risk insurance and the transfer of productive assets are examples of
social protection policies that directly reduce the poverty of low-income food pro-
ducers and increase their access to essential services For example, in India, climate-­
risk insurance policies have been developed that cover one million farmers against
crop losses related to extreme weather events (CCAFS, 2015). This, in turn,
strengthens the beneficiaries’ capacity for production and empowers them to invest
in more creative, resilient and sustainable farming techniques as well as to engage
in economically profitable activities.
Thirdly, highly technical procedures that involve institutional coordination
across all economies in wealthy and developing countries are needed. This is
because CSA implementation is knowledge-intensive and inventive. Consequently,
building capacity is a crucial component of this process (FAO, 2017a). Institutions
serve as the organisational force for those who make decisions about what to eat and
as a means of scaling up and maintaining CSA (CCAFS, 2022). Three alternative
scales, regional, national and local, should be built to create the appropriate institu-
tional capacity. According to the Consortium of International Agriculture Research
Centers (CGIAR) Research Program on Climatic Change, Agriculture and Food
Security, national institutions often play a crucial part in the provision of knowledge
about technology and management alternatives, climate variability and predictions
and market circumstances (CCAFS, 2015). For example, these institutions, through
100 E. B. Tambe et al.

research, can harmonise scientific and traditional knowledge and make available
climate forecast to farmers and advise them on certain areas where planting could
be avoided. This approach reduces economic losses. Similarly, these organisations
may support local farmers, increase their responsiveness to climate change and pro-
mote climate-smart methods that can reduce the negative impacts of agriculture on
the environment.
Building and promoting multi-stakeholder networks, collaborations and plat-
forms is an efficient strategy to guarantee the growth of institutional capacity across
different levels (FAO, 2017a, 2019c). These may make it easier to, for instance,
create innovations for climate-smart agriculture together, produce data to support
the knowledge-intensive decision-making processes involved in CSA or put CSA
alternatives into practice on the ground.
The fourth action plan is to expand available financing. Many climate-smart agri-
culture solutions have shown positive economic returns on investment, and long-­
term advantages of CSA adaptation and mitigation strategies can be anticipated for
national economies and food security (FAO, 2017a). The development of sustain-
able agriculture in poor nations can be facilitated by climate-smart agriculture
through the mobilisation of additional financial resources from bilateral partners
and international climate financing instruments (Csaky et al., 2017; Opeyemi et al.,
2021). Climate financing and official development aid for CSA may boost public
domestic and private sector investments in CSA, particularly financial services, and
help developed nations reach their USD 100 billion climate finance objective (World
Bank, 2021b).
There are also several funds available that poor nations can access to facilitate
the implementation of CSA. These funds include the Green Climate Fund (GCF),
Global Environmental Facility (GEF), Official Development Assistance (ODA),
National Sectoral Budgets, private-orientated investment and many other sources of
finance (Chiriac et  al., 2020; FAO, 2019a). For example, REDD+, Reducing
Emissions from Deforestation and Forest Degradation, is an illustration of a finance
system for sustainable forest management, plus the sustainable management of for-
ests, and the conservation and enhancement of forest carbon stocks (IFC, 2016).
With the help of this method, poor nations may get compensation for the carbon that
forests store. REDD+ can be a useful tool for developing countries to continue stor-
ing carbon on their forested land and pursue the associated adaptation and liveli-
hood co-benefits given the costs associated with sustainable forest management and
the lost economic opportunities associated with alternative land uses, like crop or
livestock production (Negra & Wollenberg, 2011).
Similarly, the International Atomic Energy Agency (IAEA), FAO, World Bank
and African Development Bank (AfDB) are also some of the organisations driving
the implementation and funding of climate-smart agriculture (UNFCCC, 2021).
These organisations support CSA in addressing the causes and effects of climate
change by monitoring agrochemical inputs for improving food safety, developing
cutting-edge land and water management technology packages and enhancing car-
bon sequestration through innovative land-water management practices. Nuclear
technology has the potential to support climate-smart agriculture in a number of
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 101

ways, including evaluating the effects of agricultural practices on climate change;


assessing the effects of climate change on agriculture; creating technologies for
adaptation, such as induced crop mutation that are resistant to abiotic stressors,
enabling them to flourish in a variety of environmental situations; water-saving
technology solutions; resilience building to climate change; and enhancing agricul-
tural practices to support climate change mitigation.
The fifth action plan is to put the procedures into action in the field. This should
capture the main goals of CSA which seek to increase the capacity of individual
food producers and other food system stakeholders to create efficient, resilient and
sustainable food production systems and value chains in the context of reducing
greenhouse gas emissions and adapting to climate change (FAO, 2019a).
Additionally, it may entail choosing flexible and appropriate CSA possibilities and
aims at involving the peasant farmers according to their expertise, needs and priori-
ties. The actions in the field should address the numerous strategies for resource use
efficiency, strengthen synergies and design pathways to reduce and accommodate
trade-offs while seeking to achieve the goals of climate-smart agriculture. Let us
examine how the goals of CSA can be achieved in the field (farm) until the food
reaches the consumer’s table.

3.3 Application of CSA to Mitigate and Adapt to Climate


Change: Farm to Table

At the farm level, several CSA approaches such as use of cleaner agricultural equip-
ment, nutrient-rich livestock feed, the use of higher-yielding seed and animal variet-
ies and the precise, timely and well-dosed application of fertilizers and pesticides
can all increase the productivity of a system while lowering the need for external
inputs, for example, use of cleaner agricultural equipment like diesel exhaust and
generating biofuel from agricultural wastes, e.g. transforming wastes such as wheat,
rice and corn straws and palm oil mill effluent (POME) to bioenergy (Gathorne-­
Hardy, 2016; Panpatte & Jhala, 2019). These approaches will reduce greenhouse
gas emissions per hectare, improve air quality and reduce exposure of farmers to air
pollution. Similarly, by increasing the recycling of wastes and by-products as inputs
within food production systems or the larger value chain, the utilisation of such
inputs may be further decreased and extend the lifespan of our resource base (FAO,
2017b). Utilising agricultural leftovers, agro-industrial waste products like oilseed
cake or manure as fertilizer are a few examples.
Climate-smart agriculture can address the effects of climate change by creating
new varieties of seeds that are tolerant to heat, accommodate pest diversity, salinity
and resistant to floods and droughts (Amin et al., 2015). Such agricultural technolo-
gies conserve resources, sustain yield and reduce emissions. Similarly, management
of pests and diseases can be a key tactic for increasing the effectiveness of resource
utilisation. Healthy organisms, including both plants and animals, may make better
102 E. B. Tambe et al.

use of agricultural inputs like fertilizers, cattle feed and fish feed, which raise the
total productivity of the corresponding production system. Additionally, pest and
disease management lowers the possibility of agricultural or animal losses, which
would increase output.
Through effective water management, such as use of drip irrigation or energy-­
efficient irrigation on agricultural lands, especially in water-stressed regions, the
amount of fossil fuel burnt for the agricultural activity per hectare is reduced
(Caldera et  al., 2021; Klauser, 2021). Similarly, strengthening capacity for using
green energy or energy-efficient mechanisation also reduces emission. Improved
management of organic wastes by encouraging aerobic decomposition through the
process of composting, or incorporating soil for the duration of the off-term drain-
age, can all reduce methane (CH4) emissions caused by rice cultivation, usage of
rice types with a better harvest index, more oxidative roots and fewer sterile tillers,
as well as the use of fermented compost instead of unfermented fertilizer as biogas
slurry (Amin et al., 2015). Reduced tillage, enhanced use of manure and organic
wastes improve the integration of soil and its biodiversity. While soil tillage is
important to ensure increase in crop production, it is essential to identify crops in
specific environments that can accommodate reduced tillage and sustain yield
(Valujeva et al., 2022). Also, partial covering of the soil might be extremely impor-
tant for the build-up of carbon content in the soil.
Crop diversification, alley cropping and agroforestry can disrupt disease cycle
and improve pest control, interaction of useful soil bacteria, suppress weed develop-
ment, increase yield, strengthen carbon capture, reduce soil erosion, enhance nutri-
ent and water use efficiency and reduce farmers’ vulnerability to shocks (Barman
et al., 2022; Siarudin et al., 2021; Wolz et al., 2018). They also strengthen land-use
efficiency and enhance soil quality. Expanding into different crops or animal spe-
cies and utilising integrated agricultural techniques like crop-livestock (silvopasto-
ral) systems, agroforestry or mixed crop-aquaculture-livestock systems are all
examples of diversification. This makes it possible to spread risks among various
agricultural operations that are exposed to climate pressures and weather extremes
to varying degrees, hence decreasing the revenue volatility of food producers. Also,
creating a favourable microclimate for crops growing beneath shade trees in agro-
forestry systems, for instance, can reduce heat stress and boost the resilience of the
entire production system (FAO, 2017a).
Efficient soil management through the application of appropriate techniques to
accommodate soil organic carbon, reduce use of chemical fertilizer and enhance the
use of livestock excrement (organic fertilizer) can achieve reduction in the input of
resources, curb emissions per hectare of food produced, increase crop yield and
reduce farmers’ resilience to shocks (Kitamura et al., 2021; Owoade, 2020; Tuğrul,
2019). For example, organic fertilizers are preferred to chemical fertilizers because
they improve soil texture and aeration, facilitate soil-water retention, stimulate the
development of healthy roots and impede processes that lead to nitrous oxide emis-
sions (Lazcano et al., 2021). Similarly, strengthening ground cover through reduc-
ing tillage and crop rotation in order to reduce erosion and land degradation is
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 103

interconnected to processes that increase crop yield, enhance carbon capture and
reduce emission of greenhouse gases.
From harvest to table, reducing food losses can be achieved through creation of
awareness, policy formulation, capacity development, incentives and redesign of
the incentive mechanism; enhancement of sustainable consumption and production;
strengthening of infrastructures to reduce spoilage/deaths in high crop/livestock
production season; increasing linkages along the food chain; strengthening of food
handling, packaging and logistic; and providing financing mechanism to farmers
(Igeta & Nakamura, 2022; Lee & Jung, 2017; Pakravan-Charvadeh & Flora, 2022).
The inevitable food wastes generated at the table or during food processing can be
transformed to boost energy security and reduce methane emissions, while the resi-
due can be used to improve soil nutrient, boost carbon capture and increase produc-
tivity (Panpatte & Jhala, 2019).
These climate-smart approaches that ensure enhanced productivity, reduced
emissions and enhanced farmers’ resilience are summarised diagrammatically
(Fig. 6.1). These approaches can be efficiently achieved in a system where there is
provision of funds; appropriate programmes and policies that are characterised with
equality, equity and justice; infrastructures; local and national institutions; interna-
tional cooperation; and goodwill.

3.4 Climate-Smart Investments, Prospects and Challenges

Globally, CSA technology is developing with support from many international


organisations in order to accomplish the triple win of increased production, resil-
ience and greenhouse gas reduction. Over US$2.5 billion has been invested in these
three areas of the CSA’s strategic goals, with the potential to benefit over 80 million
people in nations including Zimbabwe, Bangladesh, Zambia, Mali, Congo,
Morocco, Lesotho, Burkina Faso, Ghana and Cote d’Ivoire (World Bank, 2021a).
Unfortunately, these are not the only countries with climate-sensitive livelihoods.
A number of World Bank-funded initiatives in China promote institutions and
practices for resilient and low-emission agriculture, totalling US$755 million.
Through enhanced water usage efficiency on 44,000 hectares of farmland and inno-
vative technology that have improved soil conditions and increased output of rice by
12% and maize by 9%, one project has contributed to the growth of climate-smart
agriculture (World Bank, 2021b). Through this programme, more than 29,000
farmer cooperatives have reported greater revenues and increased climate resilience.
Another project that is now finished has enhanced the soil carbon sink by 71,683
tonnes of CO2 and decreased greenhouse gas emissions by 23,732 tonnes of CO2
equivalent. A project in Bangladesh seeks to strengthen the adaptability of livestock
farmers by enhancing animal health and addressing climate mitigation by enhanc-
ing emission intensity and improving production efficiency (FAO, 2014). This
includes enhancements to feeding methods, animal health, breeding, manure and
104 E. B. Tambe et al.

17
These can be achieved in a system where there is provision of funds,
appropriate policies and programs,infrastructures,local and national

Increased productivity 19 Reduced emissions 19 Enhanced resilience


18 12
11 15
10 1
institutions,international cooperation and goodwill.

3 14
9
16
13 15 14
4 2 6 8
5 7
Crop diversification Harvesting
At Farm
Drip irrigation Transportation
farm to
Renewable energy use Storage/packaging
level table
Efficient nutrient application Consumption
16
Waste management

16
15

Fig. 6.1  Agricultural changes necessary to achieve the triple win of CSA
1. Increase farmers’ income and enhance resilience
2. Strengthen soil quality and increase carbon capture
3. Enhance soil nutrient and increase productivity
4. Provide the necessary water needs and increase productivity
5. Strengthen efficiency in resource utilisation and reduce emission
6. Reduce emission per hectare
7. Reduce energy utilisation in the production chain
8. Transforming agricultural wastes to bioenergy curbs emissions
9. Increase production using organic nutrient
10. Reduce energy utilisation for production by reducing food losses associated with
harvesting
11. Appropriate timing in harvesting increases productivity
12. Timely harvesting reduces unnecessary inputs, maximises resource use and reduces
emissions
13. Strengthening transportation facilities reduced food spoilage and increase productivity
14. Reducing food losses associated with transportation assists in reducing emissions related
to production and enhance resilience
15. Improved infrastructure in storage/packaging reduces food losses and increases produc-
tion, reduces emissions associated with production and enhances resilience
16. Sustainable consumption policies reduce food wastes and reduce emissions, strengthen
environmental quality and increase productivity and enhance farmers’ resilience
17. Increased productivity enhances farmers’ income and resilience to shocks
18. Reduced emissions strengthens environmental health and increases productivity
19. Increase in sustainable production reduces emissions, guarantees farmers’ health and
enhances resilience to shocks

waste management, and low-emission technologies for tasks like milk chilling and
transport.
Brazilian researchers investigated methods for agricultural expansions to support
low-carbon agriculture while increasing private profitability as part of the sustain-
able production in areas previously converted to Agricultural Use Project (ABC
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 105

Cerrado) (FAO, 2019a). A total of 20,025 direct beneficiaries (20% female) received
technical support and training from the initiative between 2014 and 2019. These
people comprised farmers and their families, attendees at field days and associates
who were helping 378,513 hectares of land be managed sustainably. Over the fol-
lowing 10 years, it is predicted that these activities will help sequester 7.4 million
tonnes of CO2 equivalent. The Colombia Mainstreaming Sustainable Cattle
Ranching Project showing those silvopastoral systems (SPS), when combined with
additional tools for landscape management, technical support and incentives, may
result in notable successes for both farmers and the environment (World Bank,
2019). Participating farmers converted 38,390 hectares of pastureland to SPS
throughout the course of the project’s 10 years (2010–2020). Milk productivity
improved by roughly 25% compared to producing regions without SPS, milk pro-
duction costs dropped by 9% per litre, animal stocking rate rose by 26%, and farmer
revenue increased by as much as $523 per hectare annually.
As a consequence of the Mexico Sustainable Rural Development initiative, 1842
agribusinesses implemented 2286 eco-friendly technologies, including sustainable
waste management, renewable energy and energy-efficient technology (World
Bank, 2019). Similarly, in Uzbekistan, the World Bank is collaborating with the
government to support a transition away from cotton and wheat monoculture toward
a more resilient agricultural system that incorporates horticulture and uses climate-­
smart techniques that enhance soil health and lessen land degradation (FAO, 2019a).
Through the distribution of improved, drought-tolerant seeds, more effective irriga-
tion and increased use of forestry for farming and conservation agriculture prac-
tices, a bank-supported project in Niger that is specifically created to deliver
climate-smart agriculture seeks to assist 500,000 farmers and pastoralists in 44
communes (Nkonya et  al., 2018; World Bank, 2018b). The project has helped
336,518 farmers manage their land more sustainably and changed the farming tech-
niques on 79,938 hectares so far.
Improving water usage productivity in irrigated agriculture is the primary goal of
the Pakistan Punjab Irrigated Agriculture Productivity Improvement Program
Project (Pasha, 2015; Rasool & Hassan, 2017). The initiative helps to raise living
standards, agricultural productivity, employment and incomes while also having a
good impact on the environment. As of 2019, 23,500 hectares of high-efficiency
irrigation systems had been installed, and work on installing systems covering an
additional 3677 hectares was ongoing. Additionally, 11,916 watercourses had been
improved, and work on improving another 1220 was ongoing. Moreover, 5000 laser
land-levelling units had also been deployed, and 621 ponds had been built. The
initiative has produced more than 15,000 full-time employments; improved water
management benefits 5.7 million acres of agriculture and directly benefits 500,000
agricultural households.
The Climate-Smart Agriculture Project’s goal in Kenya is to help smallholder
farming and pastoral communities become more resilient to the dangers associated
with climate change (FAO, 2019a). In order to do this, climate-smart agriculture
practices must be expanded, agricultural seed systems must be strengthened, and
agro-meteorological, market, climatic and consulting services must be supported.
106 E. B. Tambe et al.

Beginning 2015, a programme funded by the World Bank has been assisting pasto-
ralists in the Sahel, specifically in Burkina Faso, Chad, Mali, Mauritania, Niger, and
Senegal, to embrace climate-smart agriculture. Initiatives to enhance rangeland
management, animal health and animal rearing are increasing production and resil-
ience and lowering emissions. Similarly, the bank is boosting CSA in Malawi by
helping farmers be more resilient to recurring and severe droughts and by raising
soil health for greater agricultural output and the adaptation and mitigation of cli-
mate change (CIAT, ICRISAT, BFS/USAID, 2020). A variety of CSA methods have
been followed by almost 140,000 farmers, and roughly 28,000 hectares of land now
have better soil health. The Maharashtra Project for Climate Resilient Agriculture,
one of the biggest CSA projects the bank has funded to date at US$420 million, is
anticipated to provide US$386 million in climate change benefits (World Bank,
2021b). As of June 2020, 56,602 hectares of land have benefited from enhanced
irrigation and drainage technology, and 309,800 project beneficiaries have embraced
climate-smart agricultural practices.
While the global climate is likely to keep changing and the necessity to keep
employing technology and genetic diversity in agriculture (Begna, 2022), invasive
species and new diseases among others are issues that will continue to face man-
kind. Similarly, with no specific formula to climate change adaptation and mitiga-
tion across regions, improving knowledge on microclimate changes that can sustain
local food production and meet the needs of societies remains frontiers of research.
Again, one of the tenets to achieve CSA is that it takes into cognizance new funding
opportunities and reduce the deficiency in investments. Despite the availability of
these funds, technologies and registered successes in CSA, the resources needed to
access these climate-smart approaches and measuring the efficiency of an approach
are quite challenging in some regions of the world, especially to those who are
dependent on climate-sensitive livelihoods (CIAT & World Bank, 2018; FAO &
ICRISAT, 2019; Jellason et al., 2018). The concern of accessing funds and technol-
ogy are made worse when violent conflicts undermine local and international coop-
eration and knowledge sharing (Fang et al., 2020). Even when these technologies
and funds are made available, it is essential to consider gender parity and equity if
the goal is to seek a productive and resilient society (Fapojuwo et  al., 2018).
Ensuring that everyone have access to this knowledge and funds, especially with the
prevalence of subsistence farming in developing countries, is an issue that needs
consistent examination in our agenda for achieving sustainable development.

4 Conclusions

The impacts of climate change and alterations on agriculture and the socio-­economic
system as a whole are growing worrisome. Following the complex nature of this
challenge, mankind has only an exit route  – designing pathways to live with the
changing climate. These pathways have been evolving, and in the agricultural sec-
tor, CSA is the contemporary approach to mitigate and adapt to climate change.
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 107

This is aligned to its scope of embracing all existing agricultural techniques and also
characterised with innovations such as new funding opportunities and reducing the
deficiency in investments, with a view of achieving increased productivity, enhanced
resilience and reduced emissions. Achieving this triple win requires a comprehen-
sive and coordinated strategy from all parties involved in the agricultural chain,
ensuring no one is left behind. While the global climate continues to keep rising,
there are variations across regions with different levels of impacts on agriculture.
Consequently, there is no unique formula to address the challenges across the globe.
Thus, consistent consideration of local priorities, enhanced synergies and reduced
trade-offs remain issues at the frontiers of research that needs a multidisciplinary
approach.

References

Abegunde, V. O., & Obi, A. (2022). The role and perspective of climate smart agriculture in Africa:
A scientific review. Sustainability, 14(4), 1–15. https://doi.org/10.3390/su14042317
Abhilash, A. R., Arti, K., Ram, N. S., & Kavita, K. (2021). Climate-smart agriculture: An integrated
approach for attaining agricultural sustainability. In V.  K. H.  Mallappa & M.  Shirur (Eds.),
Climate change and resilient food system. https://doi.org/10.1007/978-­981033-­4538-­6-­5
Akanwa, A. O., Mba, H. C., Ogbuene, E. B., Nwachukwu, M. U., & Anukwonke, C. C. (2019).
Potential of agroforestry and environmental greening for climate change minimization. In
R.  Abhishek et  al. (Eds.), Climate change and agroforestry system (p.  389). International
Standard, CRC-Apple Academic Press and Taylor & Francis.
Akinnagbe, O.  M., & Irohibe, I.  J. (2014). Agricultural adaptation strategies to climate change
impacts in Africa: A review. Bangladesh Journal of Agricultural Research, 39(3), 407–418.
https://doi.org/10.3329/bjar.v39i3.21984
Ali, N., Hu, X., & Hussain, J. (2020). The dependency of rural livelihood on forest resources in
Northern Pakistan’s Chaprote Valley. Global Ecology and Conservation, 22, 1–11. https://doi.
org/10.1016/j.gecco.2020.e01001
Amin, A., Mubeen, M., Hammad, H., & Nasim, W. (2015). Climate smart agriculture: An approach
for sustainable food security. Agricultural Research Communication, 2(3), 13–21.
Anabaraonye, B., Okafor, J.  C., Ewa, B.  O., & Anukwonke, C.  C. (2021). The impacts of cli-
mate change on soil fertility in Nigeria. In D. K. Choudhary, A. Mishra, & A. Varma (Eds.),
Climate change and the microbiome. Soil biology (Vol. 63, pp. 606–621). Springer. https://doi.
org/10.1007/978-­3-­030-­76863-­8_31
Ancha, P. U., Shomkegh, S. A., & Onuche, P. (2019). Contribution of Odoba forest reserve to live-
lihoods of the rural people in Ogbadibo Local Government Area, Benue State, Nigeria. Asian
Journal of Environment and Ecology, 9(1), 1–12.
Anuga, S. W., Gordon, C., Boon, E., & Surugu, J. M.-I. (2019). Determinants of climate smart agri-
culture (CSA) adoption among smallholder food crop farmers in the Techiman Municipality,
Ghana. Ghana Journal of Geography, 11(1), 124–139. https://doi.org/10.4314/gjg.v11i1.8
Anukwonke, C.  C., Tambe, E.  B., Nwafor, D.  C., & Malik, K.  T. (2022). Climate change and
interconnected risks to sustainable development. In S. A. Bandh (Ed.), Climate change: The
social and scientific construct (pp.  71–78). Springer International Publishing. https://doi.
org/10.1007/978-­3-­030-­86290-­9_5
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
108 E. B. Tambe et al.

Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting circular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Barman, A., Saha, P., Patel, S., & Bera, A. (2022). Crop diversification an effective strategy for
sustainable agriculture development. In V. S. Meena, M. Choudhary, R. P. Yadav, & S. Kumari
(Eds.), Sustainable crop production (pp.  1–18). IntechOpen. https://doi.org/10.5772/
intechopen.102635
Barthelmie, R. J. (2022). Impact of dietary meat and animal products on GHG footprints: The UK
and the US. Climate, 10(3), 1–18. https://doi.org/10.3390/cli10030043
Begna, T. (2022). Effect of climate change on crop genetic diversity and productivity. International
Journal of Novel Research in Civil Structural and Earth Sciences, 9(1), 1–10.
Brack, D. (2019). Forests and climate change. United Nations Forum on Forests.
Caldera, U., Sadiqa, A., Gulagi, A., & Breyer, C. (2021). Irrigation efficiency and renewable
energy powered desalination as key components of Pakistan’s water management strategy.
Smart Energy, 4. https://doi.org/10.1016/j.segy.2021.100052
CCAFS. (2015). Six steps to success. CGIAR Research Program on Climate Change, Agriculture
and Food Security (CCAFS).
CCAFS. (2022). What is climate smart agriculture. Research program on climate change,
agriculture and food security. Wageningen University and CCAFS. https://ccafs.cgiar.org/
climate-­smart-­agriculture
Celeridad, R. (2018). Contextual and universal: Scaling context-specific climate-smart agricul-
ture. World Agroforestry Centre. https://ccafs.cgiar.org/blog/contextual-­and-­universal-­scaling-­
context-­specific-­climate-­smart-­agriculture#.XLYjytJKhdi
Chiriac, D., Naran, B., & Falconer, A. (2020). Examining the climate finance gap for small-scale
agriculture. Climate policy initiative and IFAD.
Crippa, M., Solazzo, E., Guizzardi, D., Monforti-Ferrario, F., Tubiello, F. N. & Leip, P. (2021).
Food systems are responsible for a third of global anthropogenic GHG emissions. Nature Food,
2, 198–209. https://doi.org/10.1038/s43016-021-00225-9
Csaky, E., Frei-Oldenburg, A., Hess, U., Kuhn, S., Miller, C., Varangis, P., & Perry, D. (2017).
Climate smart financing for rural MSMEs: Enabling policy frameworks. G20 Global
Partnership for Financial Inclusion (GPFI).
European Commission. (2013). The impact of EU consumption on deforestation: Comprehensive
analysis of the impact of EU consumption on deforestation. European Commission.
Fang, X., Kothari, S., McLoughlin, C., & Yenice, M. (2020). The economic consequences of con-
flict in sub-Saharan Africa. International Monetary Fund (International Myeloma Foundation).
Fapojuwo, O.  E., Ogunnaike, M.  G., Shittu, A.  M., Kehinde, M.  O., & Oyawole, F.  P. (2018).
Gender gaps and adoption of climate smart practices among cereal farm households in Nigeria.
Nigerian Journal of Agricultural Economics (NJAE), 8(1), 38–49.
Food and Agriculture Organization. (2009). Global agriculture towards 2050. Food and Agriculture
Organization.
Food and Agriculture Organization. (2010). Climate-smart agriculture: Policies, practices
and financing for food security, adaptation and mitigation. Report prepared for The Hague
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 109

Conference on Agriculture, Food Security, and Climate Change October 31 to 5 November


2010. Food and Agriculture Organization.
Food and Agriculture Organization. (2014). Food wastage footprint. Full cost accounting. Food
and Agriculture Organization.
Food and Agriculture Organization. (2015). Food wastage footprint and climate change. Food and
Agriculture Organization.
Food and Agriculture Organization. (2016). Food and agriculture: Key to achieving the 2030
agenda for sustainable development. Food and Agriculture Organization. http://www.fao.
org/3/a-­i5499e.pdf
Food and Agriculture Organization. (2017a). The future of food and agriculture – Trends and chal-
lenges. Food and Agriculture Organization.
Food and Agriculture Organization. (2017b). Climate-smart agriculture sourcebook (2nd ed.). Food
and Agriculture Organization. http://www.fao.org/climate-­smart-­agriculture-­sourcebook/en
FOA (2018a). Climate-smart agriculture case studies 2018. Successful approaches from different
regions. Rome. www.fao.org/3/CA2386EN/ca2386en.pdf
FAO (2018b). Northeastern Nigeria- situation report October 2018.
Food and Agriculture Organization. (2019a). Climate-smart agriculture and the sustainable devel-
opment goals: Mapping interlinkages, synergies and trade-offs and guidelines for integrated
implementation. Food and Agriculture Organization.
Food and Agriculture Organization. (2019b). Social protection: Working towards univer-
sal coverage of social protection. Food and Agriculture Organization. http://www.fao.org/
social-­protection/overview/en/
Food and Agriculture Organization. (2019c). Guidelines for the design and implementation of
monitoring and evaluation systems for climate-smart agriculture. Food and Agriculture
Organization.
Food and Agriculture Organization. (2021). Climate smart agriculture sourcebook. Food and
Agriculture Organization. http://www.fao.org/climate-­smart-­agriculture-­sourcebook
Food and Agriculture Organization, & International Crops Research Institute for the Semi-Arid
Tropics. (2019). Climate-smart agriculture in the Adamawa State of Nigeria (CSA country
profiles for Africa series). International Center for Tropical Agriculture (Inter-American Center
of Tax Administrators), International Crops Research Institute for the Semi-Arid Tropics
(ICRISAT), Food and Agriculture Organization of the United Nations (FAO).
Gathorne-Hardy, A. (2016). The sustainability of changes in agricultural technology: The carbon,
economic and labour implications of mechanisation and synthetic fertiliser use. Ambio. Royal
Swedish Academy of Sciences,, 45(8), 885–894. https://doi.org/10.1007/s13280-­016-­0786-­5
Igeta, J., & Nakamura, H. (2022). Business incentive to reduce food losses in Japan. Sustainability,
14(4), 1–20. https://doi.org/10.3390/su14042266
Inter-American Center of Tax Administrators, & World Bank. (2018). Climate-smart agriculture in
Malawi (CSA country profiles for Africa series). International Center for Tropical Agriculture
(Inter-American Center of Tax Administrators).
Inter-American Center of Tax Administrators, International Crops Research Institute for the Semi-­
Arid Tropics, & BFS/USAID. (2020). Climate-smart agriculture in Niger (CSA country pro-
files for Africa series). International Center for Tropical Agriculture (Inter-American Center
of Tax Administrators), Bureau for Food Security, United States Agency for International
Development (BFS/USAID).
International Finance Corporation. (2016). REDD market overview. International Finance
Corporation.
IPCC. (2021). Summary for policymakers. In V. Masson-Delmotte et al. (Eds.), Climate change:
The physical science basis. Contribution of working group I to the sixth assessment report of
the intergovernmental panel on climate change (pp. 1–40). IPCC.
Jellason, N.  P., Conway, J., & Baines, R.  N. (2018). Understanding impacts and barriers to
adoption of climate-smart agriculture (CSA) practices in North-Western Nigerian drylands.
110 E. B. Tambe et al.

Journal of Agricultural Education and Extension, 27(206), 1–18. https://doi.org/10.108


0/1389224X.2020.1793787
Kitamura, R., Sugiyama, C., Yasuda, K., Nagatake, A., Yuan, Y., Du, J., Yamaki, N., Taira, K.,
Kawai, M., & Hatano, R. (2021). Effects of three types of organic fertilizers on greenhouse
gas emissions in a grassland on Andosolin Southern Hokkaido, Japan. Frontiers in Sustainable
Food Systems, 5, 649613. https://doi.org/10.3389/fsufs.2021.649613
Klauser, D. (2021). Climate-smart, resilient agriculture – Improving smallholders’ resilience, miti-
gation and profitability in all we do. Syngenta Foundation for Sustainable Agriculture.
Łaba, S., Cacak-Pietrzak, G., Łaba, R., Sułek, A., & Szczepański, K. (2022). Food losses in con-
sumer cereal production in Poland in the context of food security and environmental impact.
Agriculture, 12(5), 1–17. https://doi.org/10.3390/agriculture12050665
Lazcano, C., Zhu-Barker, X., & Decock, C. (2021). Effects of organic fertilizers on the soil micro-
organisms responsible for N2O emissions: A review. Microorganisms, 9(983), 1–18. https://
doi.org/10.3390/microorganisms9050983
Leakey, R. B. (2019). A holistic approach to sustainable agriculture: Trees, science and global
society. Burleigh Dodds Science Publishing. https://doi.org/10.19103/AS.2018.0041.25
Lee, S., & Jung, K. (2017). Exploring effective incentive design to deduce food waste: A natural
experiment of policy change from community based charge to RFID based weight charge.
Sustainability, 9(11), 1–17. https://doi.org/10.3390/su9112046
Lewis, J., & Rudnick, J. (2019). The policy enabling environment for climate smart agriculture:
A case study of California. Frontiers in Sustainable Food Systems, 3(31), 1–12. https://doi.
org/10.3389/fsufs.2019.00031
Liu, J., Hull, V., Godfray, H. C. J., Tilman, D., Gleick, P., Hoff, H., Pahl-Wostl, C., Xu, Z., Chung,
M. G., Sun, J., & Li, S. (2018). Nexus approaches to global sustainable development. Nature
Sustainability, 1(9), 466–476. https://doi.org/10.1038/s41893-­018-­0135-­8
Lynch, J., Cain, M., Frame, D., & Pierrehumbert, R. (2021). Agriculture’s contribution to cli-
mate change and role in mitigation is distinct from predominantly fossil CO2-emitting sectors.
Frontiers in Sustainable Food Systems, 4, 1–9. https://doi.org/10.3389/fsufs.2020.518039
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In S. A. Bandh (Ed.),
Climate change: The social and scientific construct. Springer Nature Publishing. https://link.
springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Matteoli, F., Schnetze, J., & Jacobs, H. (2021). Climate-smart agriculture (CSA): An integrated
approach for climate change management in the agriculture sector. In J. M. Luetz & D. Ayal
(Eds.), Handbook of climate change management (pp.  409–437). Springer. https://doi.
org/10.1007/978-­3-­030-­57281-­5_148
Mrowczyńska-Kamińska, A., Bajan, B., Pawłowski, K. P., Genstwa, N., & Zmyślona, J. (2021).
Greenhouse gas emissions intensity of food production systems and its determinants. PLoS
One, 16(4), e0250995. https://doi.org/10.1371/journal.pone.0250995
Mumtaz, M., de Oliveira, J. A., & Ali, S. H. (2019). Climate change impacts and adaptation in agri-
cultural sector: The case of local responses in Punjab, Pakistan. In S. Hussain (Ed.), Climate
change and agriculture. https://doi.org/10.5772/intechopen.83553
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Negra, C., & Wollenberg, E. (2011). Lessons from REDD+ for agriculture. CCAFS.
Nkonya, E., Ru, Y., & Kato, E. (2018). Economics of land degradation in Niger. In Fostering
transformation and growth in Niger’s agricultural sector (pp. 1–17). Wageningen Academic
Publishers. https://doi.org/10.3920/978-­90-­8686-­873-­5_2
Obinaju, L. C., & Ikpeida, D. W. (2021). Determinants of food wastes among farming households
in Uyo Local Government Area, Akwa Ibom State, Nigeria. European Journal of Agriculture
and Forestry Research, 9(4), 17–33. https://doi.org/10.37745/ejafr.2013
Opeyemi, G., Opaluwa, H. I., Adeleke, A. O., & Ugbaje, B. (2021). Effect of climate smart agri-
cultural practices on farming households’ food security in Ika North East Local Government
6  Changes in the Agriculture Sector That Are Essential to Mitigate and Adapt… 111

Area, Delta State, Nigeria. Journal of Agriculture and Food Sciences, 19(2), 30–42. https://doi.
org/10.4314/jafs.v19i2.4
OPHI, & United Nations Development Program. (2019). Global multidimensional poverty index:
Illuminating inequalities. OPHI and UNDP.
Owoade, F. M. (2020). Soil carbon management practices, knowledge of climate change and CO2
emission of some land use types in Ogbomoso Agricultural Zone, Oyo State, Nigeria. West
African Journal of Applied Ecology, 28(1), 173–188.
Pakravan-Charvadeh, M. R., & Flora, C. (2022). Sustainable food consumption pattern with empha-
sis on socioeconomic factors to reduce food waste. International Journal of Environmental
Science and Technology, 19(10), 9929–9944. https://doi.org/10.1007/s13762-­022-­04186-­9
Panpatte, D. G., & Jhala, Y. K. (2019). Agricultural waste: A suitable source for biofuel production.
In A. A. Rastegari, A. N. Yadav, & A. Gupta (Eds.), Prospects of renewable bioprocessing in
future energy systems, biofuel and biorefinery technologies (Vol. 10, pp. 337–355). Springer
Nature Switzerland. https://doi.org/10.1007/978-­3-­030-­14463-­0_13
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Pasha, H. (2015). Growth of the provincial economies. Institute for Policy Reforms.
Poore, J., & Nemecek, T. (2018). Reducing food’s environmental impacts through producers and
consumers. Science, 360(6392), 987–992. https://doi.org/10.1126/science.aaq0216
Rasool, A., & Hassan, M. A. (2017). CPEC-prospects and challenges for agriculture in Pakistan
(Vision-2025). Faisalabad Chamber of Commerce and industry. http://fcci.com.pk/rte/
Agriculture-­Report.pdf
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. https://doi.org/10.1080/1469306
2.2015.1029865
Rosenstock, T., Lamanna, C., Namoi, N., Arslan, A., & Richards, M. (2019). What is the evidence
base for climate smart agriculture in East and Southern Africa? A systematic map: Investigating
the business of a productive, resilient and low emission future. In Climate-smart agriculture
papers (pp. 141–151). https://doi.org/10.1007/978-­3-­319-­92798-­5_12
Seberini, A. (2020). Economic, social and environmental world impacts of food waste on society
and zero waste as a global approach to their elimination. SHS Web of Conferences, 74, 1–10.
https://doi.org/10.1051/shsconf/20207403010
Siarudin, M., Rahman, S.  A., Artati, Y., Indrajaya, Y., Narulita, S., Ardha, M.  J., & Larjavaara,
M. (2021). Carbon sequestration potential of agroforestry systems in degraded landscapes in
West Java, Indonesia. Forests, 12(6), 1–13. https://doi.org/10.3390/f12060714
Tambe, E. B., Okonkwo, A. U., & Eme, L. C. (2022). Determinants of volume of POME genera-
tion in palm oil mills for planning wastewater recovery in biogas energy development. Journal
of Applied Sciences and Environmental Management, 26(3), 369–376. https://doi.org/10.4314/
jasem.v26i3.1
Tuğrul, K.  M. (2019). Soil management in sustainable agriculture. In M.  Hasanuzzaman,
M.  M. Filho, M.  Fujita, & T.  A. Nogueira (Eds.), Sustainable crop production (pp.  1–15).
IntechOpen. https://doi.org/10.5772/intechopen.88319
UNFCCC. (2021). Climate smart agriculture. https://unfccc.int/BLOG/Climate-­Smart-­
Agriculture. UNFCCC.
United Nations. (2017). The sustainable development goals report. United Nations.
Valujeva, K., Pilecka-Ulcugaceva, J., Skiste, O., Liepa, S., Lagzdins, A., & Grinfelde, I. (2022).
Soil tillage and agricultural crops affect greenhouse gas emissions from Cambic Calcisol in a
temperate climate. Acta Agriculturae Scandinavica, Section B — Soil and Plant Science, 72(1),
835–846. https://doi.org/10.1080/09064710.2022.2097123. Section B.
Wolz, K. J., Lovell, S. T., Branham, B. E., Eddy, W. C., Keeley, K., Revord, R. S., Wander, M. M.,
Yang, W. H., & DeLucia, E. H. (2018). Frontiers in alley cropping: Transformative solutions
112 E. B. Tambe et al.

for temperate agriculture. Global Change Biology, 24(3), 883–894. https://doi.org/10.1111/


gcb.13986
Word Bank (2018a). Climate-smart agriculture. http://www.worldbank.org/en/topic/
climate-smart-agriculture
World Bank (2018b). Nigeria-agricultural and livestock transformation project. https://ewsdata.
rightsindevelopment.org/files/documents/09/WB-P164509
World Bank (2019). Data Bank. https://databank.worldbank.org
World Bank. (2020). Nigeria food smart country diagnostic. World Bank.
World Bank. (2021a). World Bank facts and sheets. World Bank. https://www.worldbank.org/en/
topic/climate-­smart-­agriculture
World Bank. (2021b). Shaping a climate-smart global food system. World Bank. https://www.
worldbank.org/en/topic/agriculture/publication/shaping-­a-­climate-­smart-­global-­food-­system
Chapter 7
Adaptation and Maladaptation to Climate
Change: Farmers’ Perceptions

Vahid Karimi , Masoud Bijani , Zeynab Hallaj , Naser Valizadeh ,


Negin Fallah Haghighi , and Mandana Karimi

Abstract  Nowadays, climate change is one of the most important environmental


disasters around the world and it has dramatically negative effects on agricultural
productions and farmers’ communities. Furthermore, farmers are one of the most
vulnerable communities to the impacts of climate change. Also, most researchers
have limited their research to the biological and physical fields of environmental
crises. Thus, it doubles the need for farmers’ perception and perspective on the
social and economic effects of climate change. As a result, studying farmers’ per-
ceptions and their vulnerabilities and strategies for adapting to climate change can
improve the sustainability of this part of the community, which is the most impor-
tant player in rural areas, especially in developing countries, and farmers’ maladap-
tation can be reduced with this disaster. Methodology in this research will be a
systematic review analysis by extracting and analyzing articles published in Elsevier,
Springer, and MDPI scientific databases between 2010 and 2022. Finally, in the
field of farmers’ perception of climate change, analysis of farmers’ maladaptation
in the field of climate change and identification of adaptation strategies in technical,
social, and economic dimensions will be studied.

V. Karimi · M. Bijani (*) · Z. Hallaj


Department of Agricultural Extension and Education, College of Agriculture, Tarbiat
Modares University (TMU), Tehran, Iran
e-mail: mbijani@modares.ac.ir
N. Valizadeh
Department of Agricultural Extension and Education, School of Agriculture,
Shiraz University, Shiraz, Iran
N. Fallah Haghighi
Department of Technology Development Studies (DTDS), Iranian Research Organization for
Science and Technology (IROST), Tehran, Iran
M. Karimi
Department of Sociology, University of Victoria, Victoria, BC, Canada

© The Author(s), under exclusive license to Springer Nature 113


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_7
114 V. Karimi et al.

Keywords  Climate change · Foresight · Maladaptation · Farmers’ perception ·


Adaptation strategies

1 Introduction

More than decades of research on environmental crises have revealed to the human
society that anthropogenic programs and activities are among the most important
effective components in mitigating or intensifying the impacts of these crises on
human societies, namely, local societies (Valizadeh et  al., 2021; Karimi et  al.,
2022). Nowadays, the phenomenon of climate change, which is one of the most
important environmental crises in the current decade, has caused countries to have
a deeper view at better adaptation to this crisis at the international, national, and
local levels, because it had adverse effects on various social, economic, political,
and environmental dimensions according to various studies and international
reports (Abbasian et al., 2017; Karimi et al., 2021). For this reason, adaptation to
climate change has attracted the attention of scientific and academic societies in
different regions of the world, especially in developed countries in the past years,
and in order to better adapt to this phenomenon, the decision-makers of all coun-
tries attempt to choose adaptation strategies to mitigate the adverse impacts of
climate change (Malla et  al., 2022; Bandh et  al., 2021; Chi et  al., 2020).
Considering this issue, the terms “adaptation” and “vulnerability” are widely used
to describe responses to climate change (Smith & Brown, 2014). A brief definition
of adaptation indicates that it is making some adjustments to natural or human
systems using the beneficial opportunities or reduces potential adverse impacts in
response to new or changing environments (IPCC, 2007; Nieuwaal et al., 2009).
Thus, the created adaptation strategies can reduce the social, economic, environ-
mental, etc. vulnerability caused by climate change (Brady et al., 2019). Therefore,
it can be said that vulnerability components can be a good ground for planning
and adjusting adaptation measures against the short-term and long-term outcomes
of climate change (Salik et  al., 2015). On the other hand, if the programs and
strategies used to mitigate the adverse impacts of this phenomenon are not ade-
quately efficient in terms of effectiveness, cost, and feasibility, and it exacerbates
the adverse effects of this phenomenon for societies and increases vulnerability, it
will cause the maladaptation of societies and the reduction of resilience of societ-
ies due to the ineffectiveness of maladaptive policies and actions. As a result, this
leads into irreparable impacts such as drying of international wetlands, which is
the laststep of coping up with climate change, as well as some problems such as
migration of local communities, social conflicts, and increasing poverty, jeopar-
dizing the food security of nations, sustainable livelihood of local communities,
disruption of biodiversity, increase in migrations, disturbance in ecosystems, etc
(Chi et al., 2021; Valizadeh et al., 2022). In general, now, there is no framework
for assessing maladaptation risks (Magnan, 2014). In addition, no acceptable
assessment criteria are available to determine the effects of maladaptation over
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 115

time (Granberg & Glover, 2014). The main reason for the inaccessibility of the
above instruments is that maladaptation is always exposed to unpredictable natu-
ral and human changes. Hence, with full recognition and deep understanding of
this concept and evaluating the experiences of different countries, it is possible to
help government organizations and decision-­makers to identify suitable short-
term and long-term adaptive measures and strategies by considering maladaptive
risks in the planning stage and implement them based on the conditions of each
region (Magnan et al., 2016). The present study attempts to investigate and ana-
lyze the concepts of vulnerability, adaptation, and maladaptation to climate
change in different dimensions and concepts due to extensive research in the field
of various dimensions of the technical impacts of climate change.

2 Risk Perception of Climate Change

Many environmental researchers and theorists believe that the majority of environ-
mental crisis such as climate change and the main cause of environmental problems
are referred to human behavior and anthropogenic activities (Shiri et  al., 2011;
Lechowska, 2018). Meanwhile, behavior prediction under environmental risk con-
ditions depends on how much a person is intended to recognize the risk (Lucas &
Pabuayon, 2011; Ricart et al., 2018). Thus, the decisions of societies to adopt envi-
ronmental measures are considerably affected by cognitive aspects such as aware-
ness, perceptions, expectations, and habits (Keshavarz & Karami, 2014; Tang et al.,
2013). Among the above cognitive assumptions, individual’s risk perception plays a
crucial role in shaping natural hazard policy and response management systems
(Fallah Haghighi & Bijani, 2020; Schneiderbauer et al., 2021; Soubry et al., 2020).
In recent years, there has been increasing concerns to shift the focus on fluctuations
in risk perception and identify the causes of this problem (Kahsay et  al., 2019),
because one’s perception of risk plays an important role in his/her decision-making
process (Salehi et al., 2018). Risk perception is guiding decision-making about risk
acceptance and the central influence on behavior before, during, and after a disaster
(Marshall, 2020). Also, risk perception is a mental assessment of the probability of
a special type of incident and how we concern about its consequences (Fierros-­
González & López-Feldman, 2021). Therefore, in the current situation in which
climate changes are increasing and local communities, especially farmers, need to
adapt and cope up with these changes in order to maintain and sustain their liveli-
hood resources, it is necessary to understand the different dimensions of the risk of
agricultural activities in these conditions and the relationship between risk percep-
tions and behavioral decision-making when individuals’ choices are examined in
relation to climate change (Karimi et al., 2018). In a general definition, the percep-
tion of the climate change risk means one’s perceived probability of being exposed
to the effects of climate change and one’s assessment of how these effects are harm-
ful to things that are valuable to the actor (Shen et al., 2018). In this regard, several
studies have examined climate risk perception by farmers and livestock farmers
116 V. Karimi et al.

(Lebel et al., 2015). However, in most studies on risk, risk perception is a secondary
concept, and authors have assumed that it is clearly expressed by actors. However,
the concept of risk perception needs to be examined (Miranda Sara et  al., 2016;
Schneiderbauer et al., 2021). Also, the social nature of knowing has made risk per-
ception a necessity (Bandh et  al., 2020, 2023; Nursey-Bray et  al., 2012). On the
other hand, many studies have investigated farmers’ risk perception of climate
change, which has been performed in developing countries. Recently, farmers’ risk
perception of climate change in industrialized countries has become an important
research area (Harth, 2021; Takahashi et al., 2016). Thus, risk perception is consid-
ered to maintain a central position in the planning agenda of various countries
(Barrucand et al., 2017). Also, the scientists of social science have found that people
respond to risks based on risk perception (Bustillos Ardaya et  al., 2017). In this
regard, the perception of climate risk is an important element in the attitude of farm-
ers towards adaptation. Also, risk perception influences natural hazard policy and
management systems (Fourment et al., 2020), as risk perception has long been con-
sidered as a vital determinant of human response to environmental shocks and
changes (Frank et al., 2010). However, empirical research on adaptability has mostly
ignored psychological factors, such as risk perception and perceived adaptability
capacity in determining adaptability. Most of the review of literature has identified
the issue of resource limitations as an important determinant of adaptation (Altea,
2020; Simonsson et al., 2011). Generally, various studies on climate change have
shown that a higher perception of climate change risk is associated with a greater
tendency to adapt to climate change (Carlton et al., 2015; de Mendonca & Gullo,
2020; Ullah et al., 2018).

3 Vulnerability to Climate Change

In different studies, various definitions of vulnerability have been presented.


According to the IPCC (2007), vulnerability is the degree to which a system is
susceptible to the impacts of climate change, including climate variability and the
lack of coping up with them. Vulnerability is based on some factors such as the
extent of climate change, sensitivity, and adaptive capacity of the system being
exposed (Parry et  al., 2007). Adger (2006) believes that vulnerability often
includes exposure and sensitivity to disturbances or external stress and the rele-
vant adaptation capacity (Adger, 2006). Turner et al. (2003) defines vulnerability
as the degree to which a system, subsystem, or combination of systems leads to a
disturbance or stress in exposure to risks (Turner et  al., 2003). Plummer et  al.
(2013) consider vulnerability as a key issue that refers to “the gap between expo-
sure to physical threats to human well-being and the capacity of individuals and
communities to deal with those threats” (Plummer et  al., 2013). Butler et  al.
(2014) consider vulnerability as a manifestation of poverty characterized by lim-
ited access to savings, education, health, land, housing, food, and political empow-
erment (Butler et al., 2014). Unexpected climate events and the resulting possible
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 117

disasters are not only caused by natural events; however, it is also “the product of
social, political and economic environments” (the part being separated from the
natural environment), because these environments form the life structure of differ-
ent groups of people. In other words, damages of unexpected climatic events such
as hurricanes, droughts, and floods and gradual climate changes such as high tem-
peratures are largely caused by the social, political, and economic vulnerabilities
of people on Earth. Social structuralists (e.g., those who follow a unified frame-
work), when they consider the causes of vulnerability, adopt different approaches.
By investigating social systems, they view various causes of the outcome of an
event. They also consider the integrated framework of vulnerability with having
an external dimension along an internal dimension. The external dimension is
shown by the “exposure” of the system to climatic changes; however, the internal
dimension includes its “sensitivity” and “adaptability” to these stressful factors
(Salik et al., 2015). Thus, from the perspective of a social structuralist, vulnerabil-
ity to unexpected climatic events is associated with some factors including “expo-
sure,” “sensitivity,” and “adaptability,” Exposure refers to the nature, magnitude,
and frequency of unexpected climatic events (or the degree of pressure) that a
system encounters. Sensitivity refers to the degree to which a system (e.g., a fam-
ily) is affected by unexpected climatic events. Adaptability refers to the actual or
potential ability of a system against unexpected events and successful response to
climate stimuli (McCarthy et al., 2001). Another definition of adaptability is “the
ability to design and implement effective adaptation strategies, or to respond to
increasing risks and pressure on reducing the probability of its occurrence, or
reducing the harmful outcomes that lead to damages caused by climate risks”. In
general, in the conceptual model of vulnerability in climate change conditions,
there are three key variables (sensitivity, exposure, and adaptive capacity) that
influence vulnerability in climate change conditions as presented in Fig. 7.1 (Salik
et al., 2015).
On the other hand, the agricultural sector is considered as the largest water
user. Different scenarios of climate change indicate that water stress will increase
in the future (Karimi et  al., 2018) and it is possible that in arid and semiarid

Fig. 7.1 Conceptual
framework of vulnerability
118 V. Karimi et al.

regions, the amount of agriculture fields will decrease significantly. The great
dependence of agriculture in many of these areas on surface water resources can
exacerbate the impacts of climate change and seriously threaten the production
and livelihood of farming families. In this regard, various researches have been
presented on the vulnerability of farmers under the climate change conditions.
The first study was done by Pandey et al. (2014) on “the assessment of water vul-
nerability to climate change at the household level in two forest-dependent rural
communities and semi-urban regions of the Uttarakhand Himalaya in India”. The
results of the survey study showed that the vulnerability of rural households is
higher than that of urban households. Also, the use of natural resources and cli-
matic factors have the most impact on revealing vulnerability, access to water, and
increasing search for food, the greatest effect on sensitivity to climate changes,
and social networking has the greatest effect on adaptive capacity (Pandey et al.,
2014). The research “an integrative assessment of water vulnerability in first
nation communities in Southern Ontario, Canada”, has been done by Plummer
et  al. (2013). Document analysis, questionnaire, and semistructured interviews
were used in this study. For data collection, besides participating with the natives
in document analysis, 300 samples were collected in three locations where the
natives lived, and semistructured interviews were conducted with 30 experts. The
results showed that a comprehensive overview of water vulnerability requires
adequate understanding of the level of knowledge in specific areas (Plummer
et al., 2013). Butler et al. (2014) used the framework of adaptation pathway for the
first time in the study “the application of the adaptation pathway for rural liveli-
hood and global changes in the southern island of Indonesia.” Data collection was
performed using open-ended interviews with formal and informal leaders in
southern Indonesia. They identified a total of 20 direct and indirect components
affecting rural poverty and vulnerability. Finally, the adaptation pathway was
drawn by them (Butler et al., 2014).

4 Concepts and Dimensions of Adaptation

Adaptation is a progressive and constant process that occurs over time. For having
positive adaptation to his/her changing environment, human uses inherent and
acquired mechanisms that have biological, psychological, and social origins. The
adaptive behavior is divided into two main groups:
(1) Autonomous adaptation “a response dependent on individual characteris-
tics such as what happens in natural systems and animal environments”. (2)
Planned adaptation “which is the result of organized policies based on aware-
ness and evaluation of changing conditions” (IPCC, 2007). Hence, many adap-
tation actions also reduce some risks related to the current climate fluctuations
and can support the economic and social development programs in accordance
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 119

to capacity building against water scarcity and climate change. Adaptation


refers to the issues related to communities, the systems that support these com-
munities, and their various functions in physical, economic, and natural envi-
ronments. The meaning of this term as a framework refers to a concept that can
easily be related to all the stages and departments of crisis management.
Adaptation is the features and behaviors of the system that enhance the ability
to cope with external stresses (Brooks, 2003). Adaptation includes a response to
a shock that is caused by human beings or nature and can occur before, during,
or after the event and causes the stability or improvement of social-ecological
systems (Berrang-Ford et  al., 2011). Adjustments in socioeconomic and eco-
logical systems in response to the impacts and consequences of real or expected
climate stimuli are also called adaptation to climate change (Plummer
et al., 2013).

4.1 Psychological Adaptation

Psychological adaptation is successive adaptation to changes and creating a


relationship between self and the environment in a way that maximum self-­
improvement along with social well-being is enabled while observing external
facts; thus, adaptation does not mean adjustment with others. Adaptation means
recognizing the fact that each person must pursue his/her goals according to
cultural and social frameworks (Zhang et  al., 2018). When we believe that a
person has learnt the appropriate responses in interacting with the environment,
he is adapted. A person in a specific social situation can adapt to that situation
in different ways (Kamrul Hasan & Kumar, 2020). When a person’s physical
and psychological balance is disturbed in such a way that he/she is involved in
an unpleasant state, and he/she needs to use internal forces and external support
to create balance, and he/she succeeds in this new mechanism and solves the
problem for his/her own benefit, it is said that the adaptation process has
occurred (Kamrul Hasan & Kumar, 2020. Adaptation is a dynamic, evolving
process that involves a balance between what individuals want and what their
society accepts. In other words, adaptation is a two-way process; on the one
hand, it effectively communicates with the community, and on the other hand,
the community also provides some instruments through which the individual
actualizes his/her potential. In this interaction, the individual and the society are
changed and a relatively stable adjustment is created. Adaptation dimensions
include physical adaptation, psychological adaptation, and social adaptation.
Before achieving psychological, moral, and physical adaptations, we should
adapt socially (Zhang et al., 2018)).
120 V. Karimi et al.

4.2 Social Adaptation

As adaptation encompasses a wide range and includes dimensions such as commu-


nity, family, emotions, work, health, and marriage, some experts consider social
adaptation as the top of other dimensions. The dimensions of adaptation may change
in any period of time and in any stage of social evolution. However, there is no soci-
ety that does not face the problem of maladaptation of its people. Therefore, this
section deals with a brief description of social adaptation from the point of view of
researchers and experts. From social aspects, adaptation refers to all the strategies
that a person uses to manage stressful life situations (Marshall et  al., 2010) and
social adaptation is a reflection of human interaction with other individuals, satis-
faction with individual actions, and how to perform the roles that are likely to be
affected by personality, culture, and family relationships (Marshall et al., 2010). A
person’s characteristics (skills, attitudes, values, ​​and physical states) and the sensi-
tivity of the situations that a person encounters are among the factors that affect
adaptation to the environment and its changes, and since the person and the environ-
ment are always changing, these two factors are effective on determining adapta-
tion, satisfaction, and success as adaptation occurs between these two specific
factors (Bandh, 2022; Wong et al., 2019).
Like physical, emotional, and mental growth, social adaptation is a continuous
quantity and it gradually reaches perfection and is achieved during life naturally and
in coping with experiences. After childhood and upon entering adolescence, the
psychosocial development changes from a simple transformation to a deep and
qualitative transformation, and by using social skills, the teenager can find his/her
position among social interactions and communication with his/her peers and adults
and be accepted by society. Success in social acceptance leads to social adaptation
and may guide a person to the stage of social penetration, which is a level higher
than social acceptance, and in this stage, he/she can affect other individuals. Social
penetration is a process through which human relationships move from the level of
liking to more intimacy.

5 Concepts and Dimensions of Maladaptation

An unsuccessful adaptation may simply be considered an unsuccessful action


(UNEP, 2019). However, the given adaptation is considered maladaptive when it
causes long-term or short-term vulnerability of the target communities (Barnett &
O’Neill, 2010). Maladaptation is a concept related to adaptation and vulnerability
but not well recognized (Lukasiewicz et al., 2016). Table 7.1 presents definitions of
maladaptation on climate change. Adaptation planning is an exercise in uncertainty,
and based on incomplete information, many adaptation strategies fail. Some go fur-
ther and create conditions that actually make the situation worse (Antoci et  al.,
2019). This is called maladaptation. Apart from wasting time and money,
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 121

Table 7.1  Definitions of maladaptation in the literature


Definition Source
It is an exercise if vulnerability is increased. Burton (1997)
Adaptation includes those adaptation responses that increase the vulnerability Granberg and
to climate impacts compared to the features they have, applying on other Glover (2014)
features, and exacerbating the effects in other ways, including increase in
greenhouse gas emissions.
The potential of adaptive actions that intentionally increase vulnerability is IPCC (2001)
named maladaptation.
Maladaptation in terms of specific or general resilience: Much emphasis on Walker et al.
building successful resilience for a (specified) driver (e.g., air conditioning to (2004, 2009)
be cool on a hot day) can be undermined. Resilience to other (general) stimuli
(e.g., heat tolerance in case of power disconnection causes the air conditioner
to cut off).
Incompatibility refers to adaptive actions that do not mitigate vulnerability but UNFCCC
rather exacerbate it. (2007)
Maladaptation is defined as ordinary business development that inadvertently OECD (2009)
increases exposure and/or vulnerability to climate change by ignoring the
effects of climate change. Maladaptation can also include actions taken to
adapt to climate effects that are not successful in reducing vulnerability, but
rather increase it.
Maladaptation occurs when short-term strategies increase vulnerability in the Barnett and
long-term. O’Neill (2010,
2012, 2013)
Actions being performed to prevent or reduce climate vulnerability. A change
that adversely affects or increases the vulnerability of other systems, sectors,
or social groups.
Adaptation efforts that are failed in this way or are costly in this process. Rickards and
Howden (2012)
Maladaptation arises not only from unwanted and weakly planned adaptation IPCC (2013)
actions, but also from intentional decisions in which broader considerations
overemphasize short-term outcomes and reduce or ignore long-term threats
that include a full range of interactions resulting from planned actions.
Adaptation is a process that directly or indirectly results into increased Magnan (2014)
vulnerability to climate change and change and/or significantly mitigates
current and future adaptation capacities or opportunities.
Intervention in one place or sector may increase the vulnerability of another Mycoo (2014)
place or sector or increase the vulnerability of a target group to future climate
change.
Maladaptation refers to negative changes and actions to which the families and Yaro et al.
individuals resort in response to climatic stressful factors which is opposing to (2015)
their well-being or the entire community.
Maladaptation is defined as the result of an intentional consistency policy or Juhola et al.
direct measurement of vulnerability for purposeful or external actors or the (2016)
mitigation of the prerequisites of sustainable development with the indirect
increase of the vulnerability of society.
122 V. Karimi et al.

maladaptation is a process by which individuals become vulnerable to climate


change. Weak planning is the main cause of maladaptation, yet the diverse manifes-
tations are complex and it is difficult to identify maladaptation with certainty before
(Chi et al., 2020). However, there is now enough experience to show how maladap-
tation occurs, the contexts that may be most susceptible to such an outcome, and the
design of shortcomings in strategies that should be avoided. However, unless adap-
tation projects directly address the vulnerability stimuli, maladaptation will remain
a risk. All the definitions presented in Table 7.1 indicate that maladaptation results
from adaptation that cannot reduce vulnerability or unintentionally increase it. For
example, Jones et  al. (2015) defined maladaptation as follows: “maladaptation
occurs when short-term strategies increase vulnerability in the long term.” Gersonius
et al. (2012) point out that uncertainty makes it difficult to design adaptive measures
and decide which measures are adequate to mitigate climate risks. Adaptation mea-
sures may lead to reduced reversibility in response to uncertain changes in climate
conditions. However, climate uncertainty is not the only factor that increases the
risks of maladaptation. Adaptation involves many systems, contexts, time frames,
development processes, and actors. In addition, effectiveness of adaptation is
affected by human behavior and institutional adaptation. Adaptation has a time lag
feature and different temporal and spatial effects (Barnett & O’Neill, 2010). Thus,
the results of adaptation are mostly uncertain, which can easily lead to maladaptive
decisions. In certain cases, the overall adaptive capacity is mitigated (Eriksen &
Brown, 2011).

5.1 Infrastructural Maladaptation

Researchers believe that individuals have different abilities to adapt to natural disas-
ters and their vulnerability to natural disasters is not similar; however, people’s abil-
ity can be increased in order to deal with the impacts of climate change. One of the
ways to increase adaptability in  local communities is the proper use of existing
resources and potential opportunities among local people (Anderson & Woodrow,
2019). Evidence shows that to effectively prevent and control climate change, local
communities not only resort to technical means (Breshears et al., 2011), but also,
human societies can adjust the design and structure of their systems in such a way
as to enhance their flexibility and adaptability in coping with the increasing risks
caused by weather conditions (Grothmann & Patt, 2005). Livelihood assets with
emphasis on the access and use of water resources, food and health conditions, lit-
eracy rate, and GDP industry are among the infrastructure items that are effective on
increasing the adaptability of local communities under climate change conditions
(Mohammadi et al., 2019).
The review of literature mostly focuses on maladaptation from coastal areas that
encounter the need to protect against the impacts of sea level rise, saltwater intru-
sion, coastal storms, and other climate change consequences. Various forms of
infrastructure conservation, mangrove planting, and planned retreat to adopt the
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 123

discourse of “living with floods” have been performed around the world; however,
most of these choices have consequences. A case from Fiji indicates that seawalls
built to protect people against rising sea levels actually expose people living there at
greater risk because they ultimately avoid storm water drainage. To some extent,
seawalls and other infrastructure give people a false sense of security and encourage
them to stay in locations or continue activities that make them vulnerable to climate
change if the infrastructures are damaged. In the studied example, seawalls also
transferred vulnerability to others along the coast due to changes in sediments, cre-
ating negative environmental consequences by threatening the marine ecosystem
health (Piggott-McKellar et al., 2020; Schipper, 2020). Another study in Bangladesh
investigated these actions from a gender perspective and noted that flood control has
several negative consequences including the removal of floodplains that have been
an important source of income and food and the reduction of nutrients in flooded
soils. However, these measures eliminate even more opportunities for women com-
pared to men. When these flooded areas disappeared, poor women without land
could no longer find food and resources for selling and this reduced their livelihood
security (Sultana, 2010).

5.2 Institutional Maladaptation

Institutions, as social institutions responsible for guiding and regulating social


activities, have a crucial role in coping up with the challenge of climate change
(Dovers & Hezri, 2010). Institutions not only affect how households are influenced
by climate change, they also shape the ability of households to respond to climate
change and provide different adaptation methods and mediate the external interven-
tions in the context of adaptation (Agrawal, 2001). Rural institutions are of great
importance in the formation of adaptation and its outcomes, and whether adaptation
strategies among the rural poor will be successful or not depends on the nature of
the dominant formal and informal local institutions (Glover & Granberg, 2021).
Adaptation to climate change is very important locally and its effectiveness
depends on local institutions through which incentives are provided for structured
individual and collective actions and how these institutions function at the local
level. However, efforts to adapt to the effects of climate change depend on the suc-
cess of specific institutional arrangements because adaptation never takes place in
an institutional vacuum (Glover & Granberg, 2021). The role of institutional
arrangements at the local level is in helping rural residents for effective response to
the impacts of climate change in a comprehensive and sustainable manner, build
their resilience, and protect their livelihoods against the effects of climate change.
There are many adaptation choices that can mitigate the risks of climate change
on crops and increase agricultural production (Howden et  al., 2007). Changes in
inputs such as changing cultivars/species and replacing current cultivars or species
because of needing appropriate heat period and other crop requirements with culti-
vars or species being more resistant to temperature rise and heat and drought shocks,
124 V. Karimi et al.

changing the amount of fertilizers, retaining soil moisture (e.g., leaving crop residue
on the ground), diversifying sources of income via change in the form of integrating
agriculture with other farming activities such as livestock breeding, using species or
cultivars resistant to pests and diseases, and using weather condition forecasting are
some of the factors to reduce production risks in this regard (Sahu & Mishra, 2013).
From an institutional perspective, institutional factors determine the social, eco-
nomic, and environmental outcomes of adaptation, as well as the conditions that
occurred and the processes by which such decisions are made to some extent (Glover
& Granberg, 2021). Research on agricultural climate insurance indicates that farm-
ers with insurance change how they use their land or interact differently with the
networks they previously worked on to reduce climate risks. These changes include
focusing on insured cash crops over drought-resistant livelihood crops, hybrid plant,
or moisture retention techniques, meaning farmers become reliant on insurance.
Furthermore, without the need to weigh the risks of planting different crops season-
ally, farmers no longer involve with their previous networks; thus, reducing the total
knowledge base, social capital, and risk awareness necessary to reduce uncertainty
(Müller et al., 2017).

5.3 Behavioral Maladaptation

Adapting to climate change can reduce climate damage, but adapting to climate
change is a long-term, complex, and systematic response process that individuals,
families, and farmers are faced with some barriers to use it (Castro, 2019) and are
often affected by various adaptation obstacles, and even these adaptation obstacles
lead to the unsuccessful adaptation process to climate change (Wang et al. 2020).
These barriers not only limit the ability of farmers or individuals to identify, evalu-
ate, and manage the risk of climate change (Gunathilaka et  al., 2018), but also
reduce the efficiency of household adaptation, delay in adaptation opportunities,
increase the costs of adaptation, and impede the regulation and implementation of
climate change adaptation policies (Monirul Islam et al., 2014). Thus, the inability
of some rural households to use climate change adaptation methods is related to the
barriers to adopting efficient adaptation actions in this field (Zhang et al., 2017).
Adaptation to climate change requires changes in attitudes and behavior, which
are probably more important than physical and institutional changes. However, all
behavioral changes do not seem good. A study on how farming communities
respond to climate change in northern Ghana shows how farmers temporarily
migrate from rural areas in search of work due to insecurity caused by lack of rain-
fall. But this strategy, by diversifying incomes and reducing pressure on food
reserves, leads to labor shortages so that when agricultural conditions are good, no
adequate individuals are available to ensure a successful harvest. Thus, migration
makes agriculture more difficult and changes social structures, creating new dynam-
ics and challenges (Antwi-Agyei et al., 2018). In sum, adaptation and maladaptation
should be considered as a continuum, where outcomes ranging from an ideal shift
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 125

towards a climate-resilient pathway to irreversible high vulnerability. We have many


examples of maladaptation around the world, the present condition still continues,
and the shift between adaptation and maladaptation can be subtle and fast. Similarly,
a strategy can have positive outcomes but still lead to maladaptation (Singh
et al., 2016).

6 Conclusions

The research literature shows that climate change adversely affects agriculture, and
yet the small part of agriculture that remains is the main source of livelihood among
farmers, which is one of the most important reasons for adaptation to climate change
among farmers, and adaptation is an essential tool to reduce the sensitivities and
damages caused by climate change. Therefore, considering the regular and predict-
able changes of climate change, it is necessary to design relevant response strategies
to reduce the adverse impacts of climate change.
Inadequate capital, high cost of agricultural inputs, lack of information about
adaptation, insufficient access to credit facilities, low knowledge of adaptation
options, lack of agricultural technologies, weak and unsustainable practices, socio-
cultural barriers, institutional barriers, and shortcomings of ecological zones are
among the issues that are considered some barriers to make adaptation decisions for
farmers. The literacy of farmers, agricultural experience, access to services, credit,
and climate information are factors that increase the capacity of farmers to adapt to
climate change. Also, due to the increase in population in some rural and urban
areas, climate change exposes individuals to more natural disasters and also causes
many economic and social problems; thus, creating a supportive institutional envi-
ronment for the success of adaptation efforts is necessary to provide a correct under-
standing and response to the complex interactions caused by climate change and to
help rural development and build more resistant and adaptable environments to cli-
mate change.
The importance of the government’s policies and strategic investment plans is
such that we should be ensured about the amount of access and also the improve-
ment of farmers’ access to climate forecasting information, and this access should
be economical and farmers should access and support credit schemes to increase the
ability and flexibility of their adaptation measures in response to climate conditions.
Investment in training systems and creating off-farm employment opportunities in
rural areas can be considered as a political option to reduce the adverse impacts of
climate change. Most of the financial support used by governments can be govern-
ment bonds, tax deductible reserves, public and private partnerships, and securities
related to climate change insurance.
In many developing countries, financial budgets or government bonds are regu-
larly used to pay for damages caused by natural disasters. Also, the insurance influ-
ence among people is limited, and people cannot pay for insurance coverage
obtained via insurance markets or a national public insurance program. However,
126 V. Karimi et al.

low-income countries often use the government’s financial budget as a financial


instrument, the use of insurance products against the risks of natural disasters and
the design of financial protection strategies can be very efficient, and suitable and
this plan is possible with the aid of the World Bank.
Climate change may influence regional food security using a “bottom-up”
approach that, by minimizing the performance gap and improving the product yield,
increasing efficiency in the input water and nutrients with the cultivars, resistant
products can lead to sustainable production and finally meet the food demands of
the people under climate changes.
Therefore, it is recommended to promote cost-effective, efficient, and coherent
measures to compensate for the impacts of climate change in the region. Also, con-
tinuous training of farmers about climate change and protection and conservation of
the natural environment are suggested to enhance awareness and increase the under-
standing of farmers about ways to protect natural resources.
–– Supporting farmers in the use and development of local knowledge and combin-
ing it with adaptation mechanisms introduced for local improvement
Climate Change Adaptation System
Due to climate changes and changes in the amount and the raining season, farmers
should be familiar with early and drought-resistant crops.
Water shortage is one of the major problems under climate change conditions.
Governments should introduce national or regional irrigation technologies in order
that farmers can use rainwater.
Farmers’ use of diverse livelihood systems to mitigate the effects of climate
change shocks, diversification of the agricultural system such as the use of multipur-
pose cultivation, hybrid farming, and improving the agricultural system to modern
and high-cash crops with the support of experts and the executive body of agriculture.
Integrated systems should be established in different governments and nongov-
ernmental organizations for sustainable rural development and food security of rural
residents to reduce the vulnerability of agricultural sectors against problems caused
by climate change and enhance the adaptation capacity of farmers.

References

Abbasian, A.  R., Chizari, M., & Bijani, M. (2017). Farmers’ views on the factors inhibiting
the implementation of soil conservation practices (The Case of Koohdasht Township, Iran).
Journal of Agricultural Science and Technology (JAST), 19(4), 797–807.
Adger, W.  N. (2006). Vulnerability. Global Environmental Change, 16(3), 268–281. https://doi.
org/10.1016/j.gloenvcha.2006.02.006
Agrawal, A. (2001). Common property institutions and sustainable governance of resources. World
Development, 29(10), 1649–1672. https://doi.org/10.1016/S0305-­750X(01)00063-­8
Altea, L. (2020). Perceptions of climate change and its impacts: A comparison between farmers
and institutions in the Amazonas Region of Peru. Climate and Development, 12(2), 134–146.
https://doi.org/10.1080/17565529.2019.1605285
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 127

Anderson, M. B., & Woodrow, P. J. (2019). Rising from the ashes: Development strategies in times
of disaster. Routledge.
Antoci, A., Gori, L., Sodini, M., & Ticci, E. (2019). Maladaptation and global indetermi-
nacy. Environment and Development Economics, 24(6), 643–659. https://doi.org/10.1017/
S1355770X19000251
Antwi-Agyei, P., Dougill, A. J., Stringer, L. C., & Codjoe, S. N. A. (2018). Adaptation opportuni-
ties and maladaptive outcomes in climate vulnerability hotspots of northern Ghana. Climate
Risk Management, 19, 83–93. https://doi.org/10.1016/j.crm.2017.11.003
Bandh, S. A. (2022). Climate change: The social and scientific construct (1st ed.). Springer. https://
doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S. A., Mushtaq, B., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting circular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Barnett, J., & O’Neill, S. (2010). Maladaptation. Global Environmental Change, 20(2), 211–213.
https://doi.org/10.1016/j.gloenvcha.2009.11.004
Barnett, J., & O’Neill, S. J. (2012). Islands, resettlement and adaptation. Nature Climate Change,
2(1), 8–10. https://doi.org/10.1038/nclimate1334
Barnett, J., & O’Neill, S. J. (2013). Minimising the risk of maladaptation: A framework for analy-
sis. Climate Adaptation Futures, 87–93. https://doi.org/10.1002/9781118529577.ch7
Barrucand, M. G., Giraldo Vieira, C., & Canziani, P. O. (2017). Climate change and its impacts:
Perception and adaptation in rural areas of Manizales, Colombia. Climate and Development,
9(5), 415–427. https://doi.org/10.1080/17565529.2016.1167661
Berrang-Ford, L., Ford, J. D., & Paterson, J. (2011). Are we adapting to climate change? Global
Environmental Change, 21(1), 25–33. https://doi.org/10.1016/j.gloenvcha.2010.09.012
Brady, S. P., Bolnick, D. I., Angert, A. L., Gonzalez, A., Barrett, R. D. H., Crispo, E., Derry, A. M.,
Eckert, C. G., Fraser, D. J., Fussmann, G. F., Guichard, F., Lamy, T., McAdam, A. G., Newman,
A. E. M., Paccard, A., Rolshausen, G., Simons, A. M., & Hendry, A. P. (2019). Causes of mal-
adaptation. Evolutionary Applications, 12(7), 1229–1242. https://doi.org/10.1111/eva.12844
Breshears, D. D., López-Hoffman, L., & Graumlich, L. J. (2011). When ecosystem services crash:
Preparing for big, fast, patchy climate change. Ambio, 40(3), 256–263. https://doi.org/10.1007/
s13280-­010-­0106-­4
Brooks, N. (2003). Vulnerability, risk and adaptation: A conceptual framework. Tyndall Centre for
Climate Change Research Working Paper, 38(38), 1–16.
Burton, I. (1997). Vulnerability and adaptive response in the context of climate and climate change.
Climatic Change, 36(1/2), 185–196. https://doi.org/10.1023/A:1005334926618
Bustillos Ardaya, A. B., Evers, M., & Ribbe, L. (2017). What influences disaster risk perception?
Intervention measures, flood and landslide risk perception of the population living in flood
risk areas in Rio de Janeiro state, Brazil. International Journal of Disaster Risk Reduction, 25,
227–237. https://doi.org/10.1016/j.ijdrr.2017.09.006
Butler, J. R. A., Suadnya, W., Puspadi, K., Sutaryono, Y., Wise, R. M., Skewes, T. D., Kirono, D.,
Bohensky, E. L., Handayani, T., Habibi, P., Kisman, M., Suharto, I., Hanartani, Supartarningsih,
S., Ripaldi, A., Fachry, A., Yanuartati, Y., Abbas, G., Duggan, K., & Ash, A. (2014). Framing
the application of adaptation pathways for rural livelihoods and global change in eastern
Indonesian islands. Global Environmental Change, 28, 368–382. https://doi.org/10.1016/j.
gloenvcha.2013.12.004
128 V. Karimi et al.

Carlton, J. S., Perry-Hill, R., Huber, M., & Prokopy, L. S. (2015). The climate change consensus
extends beyond climate scientists. Environmental Research Letters, 10(9), 094025.
Castro, B. (2019). The shifting limits of drought adaptation in rural Colombia. In E. Mapedza,
D. Tsegai, M. Bruntrup, & R. Mcleman (Eds.) (pp. 77–86). Elsevier. https://doi.org/10.1016/
B978-­0-­12-­814820-­4.00006-­7.
Chi, C. F., Lu, S. Y., & Lee, J. D. (2020). Ostensibly effective adaptive measures could potentially
be maladaptations: A case study of the Jiadung coastal area, Pingtung County, Taiwan. Coastal
Management, 48(6), 643–676. https://doi.org/10.1080/08920753.2020.1803575
Chi, C. F., Lu, S. Y., Hallgren, W., Ware, D., & Tomlinson, R. (2021). Role of spatial analysis
in avoiding climate change maladaptation: A systematic review. Sustainability, 13(6), 3450.
https://doi.org/10.3390/su13063450
de Mendonca, M. B., & Gullo, F. T. (2020). Landslide risk perception survey in Angra dos Reis
(Rio de Janeiro, southeastern Brazil): A contribution to support planning of non structural mea-
sures. Land Use Policy, 91. https://doi.org/10.1016/j.landusepol.2019.104415
Dovers, S. R., & Hezri, A. A. (2010). Institutions and policy processes: The means to the ends of
adaptation. Wiley Interdisciplinary Reviews: Climate Change, 1(2), 212–231.
Eriksen, S., & Brown, K. (2011). Sustainable adaptation to climate change. Climate and
Development, 3(1), 3–6. https://doi.org/10.3763/cdev.2010.0064
Fallah Haghighi, N., & Bijani, M. (2020). A gap analysis between current and desired situa-
tion of economic factors affecting human resources development in Iran. GeoJournal, 85(4),
1175–1190. https://doi.org/10.1007/s10708-019-10017-1
Fierros-González, I., & López-Feldman, A. (2021). Farmers’ perception of climate change: A
review of the literature for Latin America. Frontiers in Environmental Science, 9, 205. https://
doi.org/10.3389/fenvs.2021.672399
Fourment, M., Ferrer, M., Barbeau, G., & Quénol, H. (2020). Local perceptions, vulnerability
and adaptive responses to climate change and variability in a winegrowing region in Uruguay.
Environmental Management, 66(4), 590–599. https://doi.org/10.1007/s00267-­020-­01330-­4
Frank, E., Eakin, H., & López-Carr, D. (2010, September). Risk perception and adaptation to
climate risk in the coffee sector of Chiapas, Mexico. In Proceedings of the conference on
international research on food security, natural resource management and rural development,
Tropentag, Zurich, Switzerland (pp. 14–16).
Gersonius, B., Morselt, T., van Nieuwenhuijzen, L., Ashley, R., & Zevenbergen, C. (2012). How the
failure to account for flexibility in the economic analysis of flood risk and coastal management
strategies can result in maladaptive decisions. Journal of Waterway, Port, Coastal, and Ocean
Engineering, 138(5), 386–393. https://doi.org/10.1061/(ASCE)WW.1943-­5460.0000142
Glover, L., & Granberg, M. (2021). The politics of maladaptation. Climate, 9(5), 69. https://doi.
org/10.3390/cli9050069
Granberg, M., & Glover, L. (2014). Adaptation and maladaptation in Australian national climate
change policy. Journal of Environmental Policy & Planning, 16(2), 147–159. https://doi.org/1
0.1080/1523908X.2013.823857
Grothmann, T., & Patt, A. (2005). Adaptive capacity and human cognition: The process of indi-
vidual adaptation to climate change. Global Environmental Change, 15(3), 199–213. https://
doi.org/10.1016/j.gloenvcha.2005.01.002
Gunathilaka, R. P. D., Smart, J. C. R., & Fleming, C. M. (2018). Adaptation to climate change in
perennial cropping systems: Options, barriers and policy implications. Environmental Science
and Policy, 82, 108–116. https://doi.org/10.1016/j.envsci.2018.01.011
Harth, N. S. (2021). Affect, (group-based) emotions, and climate change action. Current Opinion
in Psychology, 42, 140–144. https://doi.org/10.1016/j.copsyc.2021.07.018
Howden, S. M., Soussana, J. F., Tubiello, F. N., Chhetri, N., Dunlop, M., & Meinke, H. (2007).
Adapting agriculture to climate change. Proceedings of the National Academy of Sciences of the
United States of America, 104(50), 19691–19696. https://doi.org/10.1073/pnas.0701890104
IPCC. (2001). 2001: Impacts, adaptation and vulnerability. Climate change (J.  J. McCarthy,
O. F. Canziani, N. A. Leary, D. J. Dokken, & K. S. White, Eds.). Cambridge University Press.
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 129

IPCC. (2007). Climate change the physical science basis. In AGU fall meeting abstracts, 2007.
IPCC. (2013). Climate change: The physical science basis. Contribution of working group I to
the fifth assessment report of the Intergovernmental Panel on Climate Change. Cambridge
University Press.
Jones, L., Carabine, E., & Schipper, L. (2015). (Re) conceptualising maladaptation in policy and
practice: Towards an evaluative framework. Available at SSRN 2643009.
Juhola, S., Glaas, E., Linnér, B. O., & Neset, T. S. (2016). Redefining maladaptation. Environmental
Science and Policy, 55, 135–140. https://doi.org/10.1016/j.envsci.2015.09.014
Kahsay, H. T., Guta, D. D., Birhanu, B. S., & Gidey, T. G. (2019). Farmers’ perceptions of cli-
mate change trends and adaptation strategies in semiarid highlands of Eastern Tigray, Northern
Ethiopia. Advances in Meteorology, 2019, 1–13. https://doi.org/10.1155/2019/3849210
Kamrul Hasan, Md., & Kumar, L. (2020). Perceived farm-level climatic impacts on coastal agri-
cultural productivity in Bangladesh. Climatic Change, 161(4), 617–636. Springer.
Karimi, V., Karami, E., & Keshavarz, M. (2018). Climate change and agriculture: Impacts
and adaptive responses in Iran. Journal of Integrative Agriculture, 17(1), 1–15. https://doi.
org/10.1016/S2095-­3119(17)61794-­5
Karimi, V., Valizadeh, N., Karami, S., & Bijani, M. (2021). Climate change and adaptation:
Recommendations for agricultural sector. In Exploring synergies and trade-offs between
climate change and the sustainable development goals (pp.  97–118). Springer. https://doi.
org/10.1007/978-­981-­15-­7301-­9_5
Karimi, V., Valizadeh, N., Rahmani, S., Bijani, M., & Karimi, M. (2022). Beyond climate change:
Impacts, adaptation strategies, and influencing factors. In Climate change (pp.  49–70).
Springer. https://doi.org/10.1007/978-­3-­030-­86290-­9_4
Keshavarz, M., & Karami, E. (2014). Farmers’ decision-making process under drought. Journal of
Arid Environments, 108, 43–56. https://doi.org/10.1016/j.jaridenv.2014.03.006
Lebel, P., Whangchai, N., Chitmanat, C., Promya, J., & Lebel, L. (2015). Perceptions of climate-­
related risks and awareness of climate change of fish cage farmers in northern Thailand. Risk
Management, 17(1), 1–22. https://doi.org/10.1057/rm.2015.4
Lechowska, E. (2018). What determines flood risk perception? A review of factors of flood risk
perception and relations between its basic elements. Natural Hazards, 94(3), 1341–1366.
https://doi.org/10.1007/s11069-­018-­3480-­z
Lucas, M.  P., & Pabuayon, I.  M. (2011). Risk perceptions, attitudes, and influential factors of
rainfed lowland rice farmers in Ilocos Norte, Philippines. Asian Journal of Agriculture and
Development, 8 (1362-2016-107714), 61–77. https://doi.org/10.22004/ag.econ.199327
Lukasiewicz, A., Pittock, J., & Finlayson, C. M. (2016). Are we adapting to climate change? A
catchment-based adaptation assessment tool for freshwater ecosystems. Climatic Change,
138(3–4), 641–654. https://doi.org/10.1007/s10584-­016-­1755-­5
Magnan, A. (2014). Avoiding maladaptation to climate change: Towards guiding principles.
S.A.P.I.EN.S. Surveys and Perspectives Integrating Environment and Society, 7(1).
Magnan, A. K., Schipper, E. L. F., Burkett, M., Bharwani, S., Burton, I., Eriksen, S., Gemenne,
F., Schaar, J., & Ziervogel, G. (2016). Addressing the risk of maladaptation to climate change.
WIREs Climate Change, 7(5), 646–665. https://doi.org/10.1002/wcc.409
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In S. A. Bandh (Ed.),
Climate change: The social and scientific construct. Springer Nature Publishing. https://link.
springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Marshall, T. M. (2020). Risk perception and safety culture: Tools for improving the implementa-
tion of disaster risk reduction strategies. International Journal of Disaster Risk Reduction, 47,
101557. https://doi.org/10.1016/j.ijdrr.2020.101557
Marshall, D., Bridges, J. F., Hauber, B., Cameron, R., Donnalley, L., Fyie, K., & Reed Johnson,
F. (2010). Conjoint analysis applications in health—How are studies being designed and
reported? An update on current practice in the published literature between 2005 and 2008.
The Patient: Patient-Centered Outcomes Research, 3, 249–256.
130 V. Karimi et al.

McCarthy, J.  J., Canziani, O.  F., Leary, N.  A., Dokken, D.  J., & White, K.  S. (Eds.). (2001).
Climate change 2001: Impacts, adaptation, and vulnerability: Contribution of working group
II to the third assessment report of the Intergovernmental Panel on Climate Change (Vol. 2).
Cambridge University Press.
Miranda Sara, L.  M., Jameson, S., Pfeffer, K., & Baud, I. (2016). Risk perception: The social
construction of spatial knowledge around climate change-related scenarios in Lima. Habitat
International, 54, 136–149. https://doi.org/10.1016/j.habitatint.2015.12.025
Mohammadi, P., Malekian, A., Ghorbani, M., & Nazri Samani, A.  A. (2019). Investigating the
relationship between the vulnerability of communities and climate changes in Kermanshah
Province. Geography and Environmental Sustainability, 9(3), 33–47. https://doi.org/10.22126/
ges.2019.3873.1994
Monirul Islam, Md., Sallu, S., Hubacek, K., & Paavola, J. (2014) Limits and barriers to adaptation
to climate variability and change in Bangladeshi coastal fishing communities. Marine Policy,
43, 208–216. https://doi.org/10.1016/j.marpol.2013.06.007
Müller, B., Johnson, L., & Kreuer, D. (2017). Maladaptive outcomes of climate insur-
ance in agriculture. Global Environmental Change, 46, 23–33. https://doi.org/10.1016/j.
gloenvcha.2017.06.010
Mycoo, M. A. (2014). Autonomous household responses and urban governance capacity building
for climate change adaptation: Georgetown, Guyana. Urban Climate, 9, 134–154. https://doi.
org/10.1016/j.uclim.2014.07.009
Nieuwaal, Kv., Driessen, P., Spit, T., & Termeer, K. (2009). A state of the art of governance lit-
erature on adaptation to climate change. Towards a research agenda. IOP Conference Series:
Earth and Environmental Science, 6(36). https://doi.org/10.1088/1755-­1307/6/36/362019
Nursey-Bray, M., Pecl, G.  T., Frusher, S., Gardner, C., Haward, M., Hobday, A.  J., Jennings,
S., Punt, A. E., Revill, H., & van Putten, I. (2012). Communicating climate change: Climate
change risk perceptions and rock lobster fishers, Tasmania. Marine Policy, 36(3), 753–759.
https://doi.org/10.1016/j.marpol.2011.10.015
Organization for Economic Cooperation and Development (OECD). (2009). Integrating cli-
mate change adaptation into development co-operation: Policy guidance. Organization for
Economic Cooperation and Development Publishing.
Pandey, R., Kala, S., & Pandey, V. (2014). Assessing climate change vulnerability of water at
household level. Mitigation and Adaptation Strategies for Global Change, 1–15. https://doi.
org/10.1007/s11027-­014-­9556-­5
Parry, M.  L., Canziani, O., Palutikof, J., Van der Linden, P., & Hanson, C. (Eds.). (2007).
2007-impacts, adaptation and vulnerability: Working group II contribution to the fourth
assessment report of the IPCC. Climate Change. Cambridge University Press, 4.
Piggott-McKellar, A. E., Nunn, P. D., McNamara, K. E., & Sekinini, S. T. (2020). Dam (n) sea-
walls: A case of climate change maladaptation in Fiji. In Managing climate change adaptation
in the Pacific region (pp. 69–84). Springer. https://doi.org/10.1007/978-­3-­030-­40552-­6_4
Plummer, R., de Grosbois, D., Armitage, D., & de Loë, R. C. (2013). An integrative assessment
of water vulnerability in First Nation communities in Southern Ontario, Canada. Global
Environmental Change, 23(4), 749–763. https://doi.org/10.1016/j.gloenvcha.2013.03.005
Ricart, S., Olcina, J., & Rico, A.  M. (2018). Evaluating public attitudes and farmers’ beliefs
towards climate change adaptation: Awareness, perception, and populism at European level.
Land, 8(1), 4. https://doi.org/10.3390/land8010004
Rickards, L., & Howden, S.  M. (2012). Transformational adaptation: Agriculture and climate
change. Crop and Pasture Science, 63(3), 240–250. https://doi.org/10.1071/CP11172
Sahu, N. C., & Mishra, D. (2013). Analysis of perception and adaptability strategies of the farmers
to climate change in Odisha, India. APCBEE Procedia, 5, 123–127. https://doi.org/10.1016/j.
apcbee.2013.05.022
Salehi, S., Chizari, M., Sadighi, H., & Bijani, M. (2018). Assessment of agricultural groundwater
users in Iran: A cultural environmental bias. Hydrogeology Journal, 26(1), 285–295. https://
doi.org/10.1007/s10040-­017-­1634-­9
7  Adaptation and Maladaptation to Climate Change: Farmers’ Perceptions 131

Salik, K. M., Jahangir, S., Zahdi, W. Z., & Hasson, S. (2015). Climate change vulnerability and
adaptation options for the coastal communities of Pakistan. Ocean & Coastal Management,
112, 61–73. https://doi.org/10.1016/j.ocecoaman.2015.05.006
Schipper, E. L. F. (2020). Maladaptation: When adaptation to climate change goes very wrong.
One Earth, 3(4), 409–414. https://doi.org/10.1016/j.oneear.2020.09.014
Schneiderbauer, S., Fontanella Pisa, P.  F., Delves, J.  L., Pedoth, L., Rufat, S., Erschbamer, M.,
Thaler, T., Carnelli, F., & Granados-Chahin, S. (2021). Risk perception of climate change and
natural hazards in global mountain regions: A critical review. Science of the Total Environment,
784, 146957. https://doi.org/10.1016/j.scitotenv.2021.146957
Shen, R., Ye, Z.  C., Gao, J., Hou, Y.  P., & Ye, H. (2018). Climate change risk perception in
global: Correlation with petroleum and liver disease: A meta-analysis. Ecotoxicology and
Environmental Safety, 166, 453–461. https://doi.org/10.1016/j.ecoenv.2018.09.080
Shiri, Sh., Bijani, M., Chaharsoughi Amin, H., Noori, H., & Soleymanifard, A. (2011). Effectiveness
Evaluation of the Axial Plan of Wheat from Expert Supervisors’ View in Ilam Province. World
Applied Sciences Journal, 14(11), 1724–1729.
Simonsson, L., Swartling, Å. G., André, K., Wallgren, O., & Klein, R. J. T. (2011). Perceptions
of risk and limits to climate change adaptation: Case studies of two Swedish urban regions.
In Climate change adaptation in developed nations (pp.  321–334). Springer. https://doi.
org/10.1007/978-­94-­007-­0567-­8_23
Singh, C., Dorward, P., & Osbahr, H. (2016). Developing a holistic approach to the analysis of
farmer decision-making: Implications for adaptation policy and practice in developing coun-
tries. Land Use Policy, 59, 329–343. https://doi.org/10.1016/j.landusepol.2016.06.041
Smith, A. M., & Brown, M. A. (2014). Policy considerations for adapting power systems to climate
change. Electricity Journal, 27(9), 112–125. https://doi.org/10.1016/j.tej.2014.10.001
Soubry, B., Sherren, K., & Thornton, T. F. (2020). Are we taking farmers seriously? A review of
the literature on farmer perceptions and climate change, 2007–2018. Journal of Rural Studies,
74, 210–222. https://doi.org/10.1016/j.jrurstud.2019.09.005
Sultana, F. (2010). Living in hazardous waterscapes: Gendered vulnerabilities and experi-
ences of floods and disasters. Environmental Hazards, 9(1), 43–53. https://doi.org/10.3763/
ehaz.2010.SI02
Takahashi, B., Burnham, M., Terracina-Hartman, C., Sopchak, A. R., & Selfa, T. (2016). Climate
change perceptions of NY state farmers: The role of risk perceptions and adaptive capacity.
Environmental Management, 58(6), 946–957. https://doi.org/10.1007/s00267-­016-­0742-­y
Tang, J., Folmer, H., & Xue, J. (2013). Estimation of awareness and perception of water scarcity
among farmers in the Guanzhong Plain, China, by means of a structural equation model. Journal
of Environmental Management, 126, 55–62. https://doi.org/10.1016/j.jenvman.2013.03.051
Turner, B. L., Kasperson, R. E., Matson, P. A., McCarthy, J. J., Corell, R. W., Christensen, L.,
Eckley, N., Kasperson, J. X., Luers, A., Martello, M. L., Polsky, C., Pulsipher, A., & Schiller,
A. (2003). A framework for vulnerability analysis in sustainability science. Proceedings of the
National Academy of Sciences of the United States of America, 100(14), 8074–8079. https://
doi.org/10.1073/pnas.1231335100
Ullah, H., Rashid, A., Liu, G., & Hussain, M. (2018). Perceptions of mountainous people on cli-
mate change, livelihood practices and climatic shocks: A case study of Swat District, Pakistan.
Urban Climate, 26, 244–257. https://doi.org/10.1016/j.uclim.2018.10.003
UNFCCC. (2007). Climate change: Impacts, vulnerabilities and adaptation in developing coun-
tries. UNFCCC.
United Nations Environment Programme (UNEP). (2019). Frontiers 2018/2019 emerging issues of
environmental concern. United Nations environmental concern. United National Environment
Programme.
Valizadeh, N., Haji, L., Bijani, M., Haghighi, N. F., Fatemi, M., Viira, A-H., Parra-Acosta, Y. K.,
Kurban, A., & Azadi, H. (2021). Development of a scale to remove farmers’ sustainability
barriers to meteorological information in Iran. Sustainability, 13(22), 12617. https://doi.
org/10.3390/su132212617
132 V. Karimi et al.

Valizadeh, N., Karimi, V., Fooladi Heleileh, B., Hayati, D., & Bijani, M. (2022). Formulating
of small-scale farmers’ perception towards climate change in arid areas: Facilitating social
interventions for agricultural sustainability. Water and Environment Journal, 36(2), 199–213.
https://doi.org/10.1111/wej.12741
Walker, B., Holling, C.  S., Carpenter, S.  R., & Kinzig, A.  P. (2004). Resilience, adaptability
and transformability in social–ecological systems. Ecology and Society, 9(2), 5. https://doi.
org/10.5751/ES-­00650-­090205
Walker, B. H., Abel, N., Anderies, J. M., & Ryan, P. (2009). Resilience, adaptability, and trans-
formability in the Goulburn-Broken catchment, Australia. Ecology and Society, 14(1), 12.
https://doi.org/10.5751/ES-­02824-­140112
Wang, W., Zhao, X., Cao, J., Li, H., & Zhang, Q. (2020). Barriers and requirements to climate
change adaptation of mountainous rural communities in developing countries: The case of
the eastern Qinghai-Tibetan Plateau of China. Land Use Policy, 95, 1688–1698. https://doi.
org/10.1016/j.landusepol.2019.104354
Wong, M. S., Lee, C., Myers, D. J., Hwang, D., Kearns, J. A., Li, T., et al. (2019). Size-independent
peak efficiency of III-nitride micro-light-emitting-diodes using chemical treatment and side-
wall passivation. Applied Physics Express, 12(9), 097004.
Yaro, J. A., Teye, J., & Bawakyillenuo, S. (2015). Local institutions and adaptive capacity to cli-
mate change/variability in the northern savannah of Ghana. Climate and Development, 7(3),
235–245. https://doi.org/10.1080/17565529.2014.951018
Zhang, Q., Zhao, X. Y., & Wang, Y. R. (2017). Adaptation needs of farmers to climate change in all
ecological vulnerable alpine region: Take Gannan Plateau for example. Acta Ecologica Sinica,
37, 1688–1698. https://doi.org/10.5846/stxb201510112054
Zhang, H., Mu, J. E., & McCarl, B. A. (2018). Adaptation to climate change via adjustment in land
leasing: Evidence from dryland wheat farms in the US Pacific Northwest. Land Use Policy,
79, 424–432.
Chapter 8
Farmers’ Perception of Climate Change
in Climatically Vulnerable Ecosystem
of Bangladesh

Foyez Ahmed Prodhan, Muhammad Ziaul Hoque, Md. Safiul Islam Afrad,


Md. Enamul Haque, Minhaz Ahmed, Md. Humayun Kabir,
Md. Sadekur Rahman, and Naima Sultana

Abstract  Climate change affects cropping seasons, water availability, and the envi-
ronment. This also increases the prevalence of pests and diseases, which ultimately
affect crop production and food availability. Hence, understanding how agricultural
communities perceive the dynamics of climate change may increase their adaptabil-
ity to it. The purpose of this research was to assess how farmers perceive climate
change and how they think it will affect agricultural yields, as well as the impact of
socioeconomic determinants on such perceptions. The investigation took place in
four upazilas in four climatically vulnerable districts of Bangladesh such as Nachole
upazila in Chapai Nawabganj District, Roumari upazila in Kurigram District,

F. A. Prodhan · M. Z. Hoque (*)


Department of Agricultural Extension and Rural Development, Bangabandhu Sheikh Mujibur
Rahman Agricultural University, Gazipur, Bangladesh
Institute of Climate Change and Environment, Bangabandhu Sheikh Mujibur Rahman
Agricultural University, Gazipur, Bangladesh
e-mail: mziahoque.aer@bsmrau.edu.bd
M. S. I. Afrad · M. E. Haque
Department of Agricultural Extension and Rural Development, Bangabandhu Sheikh Mujibur
Rahman Agricultural University, Gazipur, Bangladesh
M. Ahmed
Department of Agroforestry and Environment, Bangabandhu Sheikh Mujibur Rahman
Agricultural University, Gazipur, Bangladesh
M. H. Kabir
Department of Soil Science, Bangabandhu Sheikh Mujibur Rahman Agricultural University,
Gazipur, Bangladesh

Md. Sadekur Rahman
Department of Agricultural Extension Education, Hajee Mohammod Danesh Science and
Technology University, Dinajpur, Bangladesh
N. Sultana
National Agricultural Training Academy, Department of Agricultural Extension, Farmgate,
Dhaka, Bangladesh

© The Author(s), under exclusive license to Springer Nature 133


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_8
134 F. A. Prodhan et al.

Dumuria upazila in Kurigram District, and Teknaf upazila in Cox’s Bazar District.
A total of 320 participants were enlisted for the research using a proportional ran-
dom sampling method. Face-to-face interviews were conducted according to a pre-
tested interview plan in order to gather primary data. The research found that in all
four study areas, the majority of respondents observed a rise in temperature, drought,
and a reduction in rainfall, but there was an increased frequency of floods in Roumari
and Dumuria upazilas over the last 10–20 years. Perceptions of climate change were
shown to be significantly influenced by socioeconomic criteria such as education,
family size, agricultural experience, and access to information sources, according to
a correlation analysis. Major consequences of climate change were evident as
increased pest and disease infestation and a drop in crop production by 40–60% as
perceived by the respondents.

Keywords  Climate change · Perception · Crop production impacts · Vulnerable


ecosystem · Bangladesh

1 Introduction

The term “climate change” is used to describe any long-term shift in weather pat-
terns, whether due to natural variation or anthropogenic activity (IPCC, 2007). The
changes are caused by a combination of diverse climatic issues, such as rainfall and
temperature, as well as an increase in greenhouse gas (GHG) emissions caused by
human activity. Some of the negative implications of climate change in poor nations
include increased frequency of drought and flood, loss of biodiversity, decline of
wild and other natural resources, a rise in the prevalence of infectious illnesses and
other threats to public health, and changed livelihood patterns (Abaje & Giwa,
2007; Hassan & Nhemachena, 2008; Hassan et  al., 2021; Hoque et  al., 2019;
Prodhan et al., 2022a). Although the intensity of the effect differs significantly by
area, climate change is anticipated to have a substantial impact on agricultural pro-
ductivity and alter crop patterns (Prodhan et al., 2022b). For the millions of impov-
erished people whose lives and livelihoods rely on agriculture in this setting, the
effect of climate change on agriculture is a matter of significant worry and relevance
(Deressa & Hassan, 2009; Gwary, 2010; Hoque et  al., 2022; Prodhan et  al.,
2021, 2022a; Malla et al., 2022; Bandh et al., 2021, 2023; Mushtaq et al., 2020;
Parray et al., 2022).
Bangladesh is one of the nations across the globe that faces the greatest threat
from climate change. Bangladesh has a lot of different natural disasters like
droughts, floods, flash floods, river bank erosion, and cyclones (Prodhan et  al.,
2020, 2022b). In addition, the nation is undergoing climate change, which can be
observed in the form of an increase in temperature, an increase in the level of the
sea, and variations in the seasons, such as the way rainfalls is changing, which
affects plants and crops (Cao et al., 2022; Chowdhury et al., 2012; Hoque et al.,
2019, 2022; Nasim et  al., 2019). Changes in the world’s climate are expected to
have a big effect on agriculture in the future. This shift in weather would have severe
8  Farmers’ Perception of Climate Change in Climatically Vulnerable Ecosystem… 135

consequences for agriculture and the environment (Prodhan et al., 2022a). Rising
temperatures are the principal climate change impact on Bangladeshi agriculture,
which lead to more evaporation and droughts, which make it hard to get enough
water for irrigation and other uses in northwest Bangladesh (Hoque et  al., 2010;
Mandal, 2008; Nasim et  al., 2019; Prodhan et  al., 2020). Many issues with soil
health that reduce agricultural output are the result of a combination of climate
change and unreasonable human activity. The loss of organic matter and soil fertil-
ity, insufficient nutrients, high salinity, excessive erosion of the soil’s top layer, and
waterlogging are only some of the problems that need to be addressed. Drought has
become a common natural occurrence in this area, where crop loss ranges from 20%
to 60% (Banglapedia, 2006). The soils of Bangladesh’s char region are severely
depleted, and organic matter content is relatively low owing to light soil textures.
Productivity of these soils cannot be increased and sustained unless arrangements
are made to rebuild soil health; changes in cropping systems establish with a view
to adapt under changing climate. The coastal area of Bangladesh having huge natu-
ral resources is becoming more vulnerable because of cyclone, high tide, and inclu-
sion of salinity limiting the livelihoods of over 36 million people (Hoque et  al.,
2019, 2022; Miah et  al., 2010; Prodhan et  al., 2022b; Ziaul Hoque et  al., 2022;
Bandh et al., 2022 Bandh, 2022a, b). To develop better strategies to assist farmers in
dealing with the issues brought on by climate change, it is necessary to understand
how farmers perceive climate change and what variables influence their view.
Farmers in this area would have to adjust to climate change in order to maintain high
crop production. However, little is known about how farmers in the study region
perceived climate change. Understanding the global climate and how it is changing
it is prerequisite to adopt an adequate step to prevent climate change. As a conse-
quence, the study made an effort to assess farmers’ perception of climate change in
the study area as well as the factors driving those beliefs. In light of the aforemen-
tioned facts, the following objectives were set to guide the research: The objectives
were to (i) identify the socioeconomic features of the respondents, (ii) determine the
perceptions of respondents on climate change and its impact on agricultural produc-
tion, and (iii) analyze how respondents’ socioeconomic status affects their percep-
tion on climate change in the study region.

2 Methodology

The study included four climatically vulnerable districts, namely, Chapai Nawabganj,
Kurigram, Khulna, and Cox’s Bazar. The research locations were chosen using a mul-
tistage random sampling process. After selecting four districts, one upazila from each
district, namely, (i) Nachole upazila of Chapai Nawabganj District, (ii) Roumari upazila
of Kurigram District, (iii) Dumuria upazila of Khulna District, and (iii) Teknaf upazila
of Cox’s Bazar District (Fig. 8.1), was selected purposively considering the type and
nature of climatic hazard and the livelihood system of the farmers. The sample size in
each upazila across all districts was determined using a proportional random sampling
method. A total of 320 farmers, 80 from each upazila, were chosen as a representative
136 F. A. Prodhan et al.

Fig. 8.1  Study area map

sample for the research using random sampling techniques. A pretested interview
schedule was employed to obtain data from respondents. Respondents were requested
to provide personal information such as their age, education, farm size, yearly income,
8  Farmers’ Perception of Climate Change in Climatically Vulnerable Ecosystem… 137

training, and mass media exposure. Farmers were asked if temperatures, rainfall,
droughts, and floods had increased, decreased, or stayed the same since their adoles-
cence as compared to the recent past. The questions were designed so that respondents
could compare situations in the recent past (less than 5  years) and the distant past
(period since their adolescence) to assess any perceived changes in length. To evaluate
the findings, the obtained data was collated and statistically examined. Quantitative
data was analyzed using descriptive statistics, and findings were presented as percent-
ages or counts to reflect farmers’ perspectives on climate change. Using Pearson’s cor-
relation, the relationship between some farmer traits and how they felt about climate
change was looked into.

3 Findings and Discussion

3.1 Features of the Respondents’ Socioeconomic Status

This section explored the socioeconomic characteristics of the study’s respondents.


The purpose was to get an idea about the population characteristics of climatically
vulnerable areas’ farmers. Table  8.1 displays the socioeconomic parameters

Table 8.1  Distribution of respondents based on their social and economic status
Measuring No. of
Character unit Categories respondents Percent Mean SD
Age Actual year Young aged (up to 98 36.75 43.25 10.54
35)
Middle aged (36–45) 129 48.38
Old (>45) 93 34.88
Education Year of Illiterate (0) 90 33.75 4.05 3.76
schooling Primary (1–5) 87 32.63
Secondary (6–10) 73 27.38
Higher secondary 70 26.25
(>10)
Family size Number Small (up to 4) 113 42.38 4.68 4.90
Medium (5–6) 119 44.63
Large (7 and above) 88 33.00
Farm size Actual Small (up to 1) 153 57.38 0.94 0.65
(ha) Medium(1.01–3.0) 98 36.75
Large (above3.0) 69 25.88
Farming No. of years Poor (up to 15) 106 39.75 10.75 6.01
experience Moderate (16–20) 137 51.38
High (above 20) 77 28.88
Exposure to Scores Low (up to 15) 87 32.63 20.86 3.23
mass media Medium (16–25) 140 52.50
High (above 25) 93 34.88
138 F. A. Prodhan et al.

including age, education, family size, farm size, agricultural experience, and inter-
action with information sources. The majority of respondents (48.38%) were of
middle age, compared to 34.88% of older respondents and 36.75% of younger
respondents.
In contrast to the 33.75% of respondents who were illiterate, the biggest percent-
age of respondents (32.63%) had a primary level of education, followed by second-
ary education (27.38%) and upper secondary education (26.25%). More than
two-fifths (42.38% and 44.63%) of the respondents falls under both small and
medium families as compared with the 33% large category. Moreover, half of the
respondents (57.38%) had small farms, while 36.75% had medium farms, and
25.88% had big farms. Among those who answered the survey, 51.38% had some
agricultural experience, 39.75% had none, and 28.88% had extensive farming expe-
rience. Most of the respondents (52.50%) were found to have medium exposure to
mass media, while 34.88% were in high and 32.63% respondents were found to
have low contact with sources of information.

3.2 Farmers’ Perception to Climate Change

The majority of respondents in the study regions saw changes in temperature, pre-
cipitation, and flood and drought conditions; therefore, they had a good understand-
ing of climate change (Figs.  8.1, 8.2, 8.3 and 8.4). Changes in perceptions of
temperature and precipitation, as well as flood and drought, were classified into four
distinct areas. Nearly all participants agreed that the shift in climate conditions over
the previous two decades was undeniable.

73.33 70.00
62.50
56.67
% Respondents

Increased
33.33
26.67 25.83 Decreased
21.67
11.67 8.33 Unchanged
10
0

Nachole Roumary Dumuria Teknaf

Fig. 8.2  Farmers’ perception on the pattern of temperature changes in four climatically vulnera-
ble areas
8  Farmers’ Perception of Climate Change in Climatically Vulnerable Ecosystem… 139

76.67
64.17
58.33
54.17
%Respondents

Increased
35.83
30.83 Decreased
20.83
15 15.00 18.33
Unchanged
8.33 10.83

Nachole Roumary Dumuria Teknaf

Fig. 8.3  Farmers’ perception on the trend of rainfall changes in four climatically vulnerable areas

79.17

50 50.00
%Respondents

45.83
40.00 Increased

27.50 29.17 Decreased


22.50 25
20.83 Unchanged
10
0

Nachole Roumary Dumuria Teknaf

Fig. 8.4  Farmers’ perception on the trend of flood changes in four climatically vulnerable areas

3.2.1 Farmers’ Perception on Temperature Changes

The result indicates that about three-fourths (73.33%) of the respondents perceived
that the temperature of Nachole upazila in Chapai Nawabganj District is increasing
with time. On the contrary only 8% of them have not noticed any change in tem-
perature. At Roumari in Kurigram District more than two-fifths (62.5%) of the
respondents observed an increase in temperature during the last two decades, while
56.67% of respondents have noticed the temperature is increasing for the last
10–20 years up to the recent time at Dumuria in Khulna District. In case of Teknaf
in Cox’s Bazar District 70% of respondents expressed their opinion that tempera-
ture is also increasing with time. It can be concluded that temperature at the four
upazilas in four climatically vulnerable districts is increasing over the past
10–20 years (Fig. 8.2).
140 F. A. Prodhan et al.

3.2.2 Farmers’ Perception on Rainfall Changes

The information contained in Fig. 8.3 revealed that more than three-fifths (76.67%)
of the respondents perceived that the rainfall of Nachole upazila in Chapai
Nawabganj District is decreasing with time. At Roumari in Kurigram District 64.7%
of the respondents observed a decrease in rainfall during the last two decades, while
58.33% of respondents have noticed rainfall is increasing for the last 10–20 years up
to the recent time at Dumuria in Khulna District. In case of Teknaf in Cox’s Bazar
District more than half (54.17%) of respondents expressed their opinion that rainfall
is also increasing with time. Results also indicate that the rainfall trend at the four
upazilas in four climatically vulnerable districts is decreasing over the past
10–20 years.

3.2.3 Farmers’ Perception on Flood

The result presented in Fig.  8.4 revealed that about fourth-fifths (79.17%) of the
respondents perceived that the flood incidence in Nachole upazila in Chapai
Nawabganj District is decreasing with time. At Roumari in Kurigram district half of
the respondents (50%) observed an increase in flood incidence during the last two
decades, while 45.83% of respondents have noticed flood incidence to be increasing
for the last 10–20 years up to the recent time at Dumuria in Khulna District. In case
of Teknaf in Cox’s Bazar District 50% of respondents expressed their opinion that
flood incidence is also increasing with time. Results also indicate that the flood
occurrence trend at Roumari and Dumuria upazilas in Kurigram and Khulna
Districts, respectively, is increasing, while it is decreasing in Nachole and Teknaf
over the past 10–20 years.

3.2.4 Farmers’ Perception on Drought

The information contained in Fig. 8.5 revealed that about fourth-fifths (79.17%) of


the respondents perceived that the drought of Nachole upazila in Chapai Nawabganj
District is increasing with time. At Roumari in Kurigram District three-fifths
(60.83%) of the respondents observed an increase in drought during the last two
decades, while half of the respondents (50%) have noticed drought is increasing for
the last 10–20 years up to the recent time at Dumuria in Khulna District. In case of
Teknaf in Cox’s Bazar District 54.17% of respondents expressed their opinion that
drought is also increasing with time. Results also indicate that drought occurring at
four upazilas in four climatically vulnerable districts is increasing over the past
10–20 years.
79.17

60.83
54.17
50
Increased
35.83
29.17 Decreased
25

%Respondents
20.83 Unchanged
14.17 14.17 16.67

Nachole Roumary Dumuria Teknaf

Fig. 8.5  Farmers’ perception on the trend of drought changes in four climatically vulnerable areas
8  Farmers’ Perception of Climate Change in Climatically Vulnerable Ecosystem…
141
142 F. A. Prodhan et al.

3.2.5 Perception of Farmers About the Impact of Climate Change

The respondents’ perceived impact of climate change on crop production is dis-


cussed in this section. The purpose was to get an idea about climate change and
climate variability issue such as temperature, flood, rainfall, drought, and salinity on
crop production and its yield.
The results presented in Table 8.2 show that most of the farmers reported that
about 11% of cultivable land at Nachole in Chapai Nawabganj District remained
uncultivated in Rabi season because of drought, while at Roumari in Kurigram
District about 8% land in Kharif-1 season remained fallow due to flood (Table 8.2).
At Dumuria and Teknaf upazilas about 18% and 24% land, respectively, remained
uncultivated in Kharif-1 season because of salinity.
The study showed that because of climatic hazards, disease and pest infestation
increased with time in Kharif-1 season (Table  8.3). It was found that at Nachole
upazila 65% of respondents perceived that severe disease/pest infestation increased
in Kharif-1 due to drought. At Roumari more than three-fifths of respondents (62%)
expressed their opinion that maximum disease/pest infestation was found in Kharif-2
because of regular flood. Salinity in the coastal area, i.e., Dumuria (57.5% of respon-
dents) and Teknaf (64.7% of respondents), was observed as the main factor of dis-
ease/pest infestation.
The substantial reduction in crop yields (40–60%) was attributed to drought,
erratic rainfall, flood, river bank erosion, and salinity (Table 8.4). From the respon-
dents’ opinion it was observed that climatic events such as drought and erratic rain-
fall cause severe yield loss (40–60%) at Nachole upazila in Chapai Nawabganj
District, while flood and erratic rainfall reduced crop yields severely at Roumari

Table 8.2  Uncultivable land due to drought, flood, and salinity problems of the respondents in
four climatically vulnerable areas
Nachole Roumari Dumuria Teknaf
Drought Drought Flood Flood Salinity Flood Salinity
Season (%) (%) (%) Drought (%) (%) Drought (%) (%)
Rabi 10.54 4.53 5.1 – 5.05 6.04 – 5.79 7.38
Kharif-1 5.75 2.78 7.5 – 4.08 17.59 – 4 24.23
Kharif-2 7.84 3.93 9.67 – 10.08 7.01 – 6.41 9.51

Table 8.3  Severity of disease/pest infestation due to drought, flood, and salinity problems at four
upazilas in climatically vulnerable districts
Nachole Roumari Dumuria Teknaf
Season Drought Flood Flood Salinity Flood Salinity
Respondents opinion (%)
Rabi 65.00 29.80 41.70 43.30 39.20 48.00
Kharif-1 74.00 51.90 38.30 57.50 35.30 64.70
Kharif-2 50.00 62.70 56.70 23.30 47.60 13.70
Table 8.4  Perceived climate change impact on crop yield
Nachole Roumari Dumuria Teknaf
Yield
Climate change and loss
climate variability issues Crops (%) Severity Yield loss Severity Yield loss Severity Yield loss Severity
Sea level rise (SLR)/ Cereal crops, 0 – 0 – 40–60 Severe 40–60 Severe
inundation Tuber crops,
Salt water intrusion/ Vegetables, 0 – 0 – >60 Very severe >60 Very severe
salinity intrusion Pulse crops,
Increase of tropical Oilseed crops, 0 – 0 – 20–40 Moderate 20–40 Moderate
cyclones Spice and fruit
crops
Storm surges 0 – 0 – <20 Low <20 Low
Tidal surge 0 0 – <20 Low 20–40 Moderate
Flood 0 – 40–60 Severe 20–40 Moderate 20–40 Moderate
Waterlogging 0 – 20–40 Moderate 20–40 Moderate 20–40 Moderate
River bank erosions 0 – 40–60 Severe 0 – 0 –
Sedimentation/carpeting 0 – 0 – 0 – 0 –
Drought 40– Severe <20 Low <20 Low <20 Low
60
Erratic rainfall/change 40– Severe 20–40 Moderate <20 Low <20 Low
and unpredictability in 60
rain pattern
Increased rainfall 0 – 0 – 0 – 0 –
Decreased rainfall 0 – 0 – 0 – 0 –
Heat waves 20– Moderate <20 Low 0 – <20 Low
8  Farmers’ Perception of Climate Change in Climatically Vulnerable Ecosystem…

40
Cold waves 20– Moderate 20–40 Moderate <20 Low <20 Low
40
143
144 F. A. Prodhan et al.

Table 8.5 Relationship between socioeconomic characteristics and their perception of


climate change
Socioeconomic Coefficient of correlation (‘r’) value for climate change
characteristic perception
Age −0.090NS
Education 0.314(**)
Family size 0.260(**)
Farm size 0.040NS
Farming experience 0.185(*)
Exposures to mass media 0.667(**)
“*” denotes significance at the 0.05 level; “**” represents significance at the 0.01 level; “NS”
indicates nonsignificant

upazila in Kurigram District. Salinity and/or salt water intrusion caused very severe
yield loss (>60%) at Dumuria and Teknaf upazila in Khulna and Cox’s Bazar
District, respectively.

3.3 Factors Affecting Farmer Perceptions to Climate Change

The analysis findings, which looked at the variables affecting farmers’ perception of
climate change, are shown in Table 8.5. It depicts the findings of an investigation
into the variables affecting farmer perceptions of climate change. The findings dem-
onstrated a positive and substantial link between farmer beliefs of climate change
and education, family size, agricultural experience, and media exposure. This sug-
gests that a farmer’s ability to recognize and adapt to the effects of climate change
increases with his/her level of education. In the same way, the longer an individual
farms, the more knowledge he/she gains about climate change. A large family size
may signify increased information exchange on climate change inside the family
among family members. The farmers who have greater exposure to the media are
perhaps better qualified to comment on whether climate change has really hap-
pened. Generally speaking, however, the findings suggest that the farmers’ socio-
economic traits affect how they perceive climate change. Farmers who have had
greater exposure to the media are apparently better prepared to comment on whether
climate change has happened. However, the findings suggest that the farmers’
socioeconomic traits impact their view of climate change.

4 Conclusions

The study’s findings indicated that the majority of farmers were middle-aged and
illiterate and had small farms with moderate exposure to the media. Rainfall pattern,
temperature, flood, and drought are examples of climate elements that they believe
8  Farmers’ Perception of Climate Change in Climatically Vulnerable Ecosystem… 145

have changed dramatically. They seem to have had a good understanding of climate
change. The vast majority of responders were of the opinion that both the tempera-
ture and drought are becoming worse, while rainfall is decreasing with time at four
upazilas in four climatically vulnerable districts, but the flood occurrence trend at
Roumari in Kurigram District and Dumuria in Khulna District is increasing, while
it is decreasing at Nachole in Chapai Nawabganj District and Teknaf in Cox’s Bazar
District over the past 10–20  years. Climate change perceptions in the research
region were heavily impacted by respondents’ socioeconomic status. Farmers’
views on climate change were profoundly affected by their educational background,
family size, agricultural experience, and exposure to mass media. Governments and
nongovernment organizations should set up workshops for extension agents and run
campaigns through the media to teach farmers about climate change, especially its
causes and effects, and how to adapt to it.

References

Abaje, I.  B., & Giwa, P.  N. (2007). Urban flooding and environmental safety: A case study of
Kafanchan in Kaduna State, Nigeria. A paper presented at the golden jubilee (50th anniver-
sary) and 49th annual conference of the Association of Nigerian Geographers (ANG) at the
university of Abuja, October 15–19.
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting circular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Banglapedia. (2006). National encyclopedia of Bangladesh. http://www.banglapedia.org/httpdocs/
HT/D_0284.HTM
Cao, D., Zhang, J., Han, J., Zhang, T., Yang, S., Wang, J., Prodhan, F. A., & Yao, F. (2022). Projected
increases in global terrestrial net primary productivity loss caused by drought under climate
change. Earth’s Future, 10(7), e2022EF002681. https://doi.org/10.1029/2022EF002681
Chowdhury, A. K. M. H. U., Haque, M. E., Hoque, M. Z., & Rokonuzzam, M. (2012). Adoption of
BRRI dhan47 in the coastal saline areas of Bangladesh. Agricultural Journal, 7(5), 286–291.
https://doi.org/10.3923/aj.2012.286.291
Deressa, T. T., & Hassan, R. M. (2009). Economic impact of climate change on crop production
in Ethiopia: Evidence from cross-sectional measures. Journal of African Economies, 18(4),
529–554. https://doi.org/10.1093/jae/ejp002
146 F. A. Prodhan et al.

Gwary, D. M. (2010). Climate change adaptation and mitigation options for improving food secu-
rity in Nigeria. A lead paper presented at 6th national conference on organic agriculture at
University of Maiduguri, Nigeria, November 22–24.
Hassan, R., & Nhemachena, C. (2008). Determinants of African farmers’ strategies for adapting
to climate change: Multinomial choice analysis. African Journal of Agricultural and Resource
Economics, 2(1), 83–104.
Hassan, T., Zhang, J., Prodhan, F. A., Pangali Sharma, T. P., & Bashir, B. (2021). Surface urban
heat islands dynamics in response to LULC and vegetation across South Asia (2000–2019).
Remote Sensing, 13(16), 3177. https://doi.org/10.3390/rs13163177
Hoque, M. Z., Haque, M. E., Afrad, M. S. I., & Hossain, M. A. (2010). Adoption of farming tech-
nology by the Charland farmers. Bangladesh Journal of Extension Education, 22(1&2), 49–55.
Hoque, M. Z., Cui, S., Xu, L., Islam, I., Tang, J., & Ding, S. (2019). Assessing agricultural livelihood
vulnerability to climate change in coastal Bangladesh. International Journal of Environmental
Research and Public Health, 16(22), 4552. https://doi.org/10.3390/ijerph16224552
Hoque, M. Z., Haque, M. E., & Islam, M. S. (2022). Mapping integrated vulnerability of coastal
agricultural livelihood to climate change in Bangladesh: Implications for spatial adapta-
tion planning. Physics and Chemistry of the Earth, Parts A/B/C, 125, 103080. https://doi.
org/10.1016/j.pce.2021.103080
IPCC. (2007). Climate change – Impacts, adaptation and vulnerability: Contribution of working
group JP IIto the fourth assessment report of the Intergovernmental Panel on Climate Change.
Cambridge University Press.
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In S. A. Bandh (Ed.),
Climate change: The social and scientific construct. Springer Nature Publishing. https://link.
springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Mandal, M. A. S. (2008). World food security: The challenges of climate change and bioenergy-­
Bangladesh perspectives. On Keynote paper presented at the World Food Day 2008 seminar
“World food security: The challenges of climate change and bioenergy”, organized by the
Ministry of Agriculture, Government of the People’s Republic of Bangladesh at the Bangladesh,
October 16, 2008. China Friendship Conference Centre.
Miah, G. M., Bari, N. M., Islam, F. S. M., Ahamed, T., & Rahman, M. A. (2010). Impacts of anthro-
pogenic activities on natural resources and food security in the coastal region of Bangladesh.
National Food Policy Capacity Strengthening Program, United States Agency for International
Development, European Commission, Food and Agriculture Organization.
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Nasim, F.  A., Hoque, M.  Z., Haque, M.  E., Islam, M.  S., Parveen, N., Chakma, S., & Afrad,
M. S. I. (2019). How does adoption of crop variety reduce the impact of drought in agriculture
and mitigate food insecurity of smallholder farmers? A case study on BUdhan1 rice variety in
Bangladesh. Asian Journal of Agricultural Extension, Economics and Sociology, 1–12. https://
doi.org/10.9734/ajaees/2019/v30i330114
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Prodhan, F.  A., Zhang, J., Bai, Y., Pangali Sharma, T.  P. P., & Koju, U.  A. (2020). Monitoring
of drought condition and risk in Bangladesh combined data from satellite and ground
meteorological observations. IEEE Access, 8, 93264–93282. https://doi.org/10.1109/
ACCESS.2020.2993025
Prodhan, F. A., Zhang, J., Yao, F., Shi, L., Pangali Sharma, T. P., Zhang, D., Cao, D., Zheng, M.,
Ahmed, N., & Mohana, H.  P. (2021). Deep learning for monitoring agricultural drought in
South Asia using remote sensing data. Remote Sensing, 13(9), 1715. https://doi.org/10.3390/
rs13091715
8  Farmers’ Perception of Climate Change in Climatically Vulnerable Ecosystem… 147

Prodhan, F. A., Zhang, J., Hasan, S. S., Pangali Sharma, T. P. P., & Mohana, H. P. (2022a). A review
of machine learning methods for drought hazard monitoring and forecasting: Current research
trends, challenges, and future research directions. Environmental Modelling and Software, 149,
105327. https://doi.org/10.1016/j.envsoft.2022.105327
Prodhan, F. A., Zhang, J., Pangali Sharma, T. P. P., Nanzad, L., Zhang, D., Seka, A. M., Ahmed, N.,
Hasan, S. S., Hoque, M. Z., & Mohana, H. P. (2022b). Projection of future drought and its impact
on simulated crop yield over South Asia using ensemble machine learning approach. Science of
the Total Environment, 807(3), 151029. https://doi.org/10.1016/j.scitotenv.2021.151029
Ziaul Hoque, M., Islam, I., Ahmed, M., Shamim Hasan, S., & Ahmed Prodhan, F. (2022). Spatio-­
temporal changes of land use land cover and ecosystem service values in coastal Bangladesh.
Egyptian Journal of Remote Sensing and Space Science, 25(1), 173–180. https://doi.
org/10.1016/j.ejrs.2022.01.008
Chapter 9
Pest and Disease Management Under
Changing Climate

Yaser Biniaz, Naser Valizadeh , Farshad Hemmati ,


and Alireza Afsharifar

Abstract  Global warming leads to climate change at certain times and in different
geographical areas. Climate change has serious consequences for agricultural pro-
duction and food security for the world’s growing population. The consequences of
climate variability such as changes in precipitation patterns, increasing tempera-
tures, and increasing wind speeds have direct impacts on the dynamics and epidem-
ics of plant diseases and pests. In addition, factors such as substitute products and
changes in the distribution and activity of vectors can indirectly affect the develop-
ment of these diseases and pests. In other words, climate change leads to changes in
the range of activity of pests and diseases, relative abundance of pathogens, preva-
lence rate, resistance to the pathogen, physiology of host–pathogen interaction, rate
of evolution, host compatibility, and effectiveness of management measures of the
damaging epidemics. In this regard, pest and disease management under climate
change was determined as the main objective of this chapter. For this purpose, a
systematic review of the literature was used in order to develop a practical model of
pest and disease management under climate change. Finally, based on the frame-
work developed, policy-makers, managers, practitioners, and farmers were pro-
vided with some practical recommendations.

Keywords  Climate change · Pest and disease management · Risk reduction ·


Disease management systems

Y. Biniaz · F. Hemmati · A. Afsharifar


Plant Virology Research Center, School of Agriculture, Shiraz University, Shiraz, Iran
N. Valizadeh (*)
Department of Agricultural Extension and Education, School of Agriculture, Shiraz
University, Shiraz, Iran
e-mail: n.valizadeh@shirazu.ac.ir

© The Author(s), under exclusive license to Springer Nature 149


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_9
150 Y. Biniaz et al.

1 Introduction

Currently, the earth is warming due to periodic cycles of climate change (Garrett
et al., 2006; Lovisolo et al., 2003). These climate changes affect the main parts of
the life of plants and organisms. Scientists believe that human activity is one of the
most important causes of temperature increase due to the production of more carbon
dioxide and other greenhouse gases (mainly methane and nitrogen oxides). During
the last one hundred years, the average temperature of the world has increased up to
0.8 °C (Kliejunas, 2010). In fact, in the second climate change report published by
the IPCC, three main conclusions were emphasized: (1) The concentration of green-
house gases has increased and will continue to exist in the atmosphere for a long
time. (2) Since the last century, due to the increase in the amount of carbon dioxide,
the climate has changed and the temperature has increased between 0.3 and
0.6 °C. (3) The use of man-made sulfur sprays, such as sulfur dioxide, which have
a negative effect on the effects of radiation and have a short life, has increased. And
(4) in climate simulation models, it is predicted that by the year 2100, the earth will
be between 1 and 3.5 °C warmer. This IPCC report became the basis for discussion
at the Kyoto conference in Japan, which was approved by more than 150 countries
(Mushtaq et  al., 2020; Bandh et  al., 2021; Bandh, 2022a, b; Bandh et  al., 2022,
2023; Malla et al., 2022; Parray et al., 2022).
Human intervention in nature, excessive consumption of fossil fuels, and changes
in the use of land (using the destruction of forests) have caused climate change.
These factors have led to the reduction of the earth’s oxygen production capacity
and, as a result, the accumulation of more greenhouse gases that remain in the atmo-
sphere for a long time. In such a situation, due to the absorption and re-reflection of
infrared rays, the earth will become warmer (Mahlman, 1997). Many models have
been reported for predicting climate change. Most of these forecasts are based on
the production of greenhouse gases. However, it is necessary to emphasize that
information for predicting climate changes should be used from more data sources.
Until 1990, there was very little information about the effects of climate change on
plant pests and diseases. For example, in 1998, a book was published about climate
change and its impact on agriculture (Rosenzweig & Hillel, 1998). Interestingly, out
of 326 pages, only two pages were related to pests and plant diseases. That is while,
a significant amount of the content of the book was related to human diseases. In
other words, the book focused more on the impacts of global warming on the
increase of human diseases.
According to the estimates of some climate change models, by the year 2100 the
average global temperature can increase by 6 °C (Norse, 2003). Of course, warming
occurs faster in higher latitudes than in lower latitudes. In other words, at higher
latitudes, weather events will be more frequent and more intense. Spatial and tem-
poral climate changes have major consequences in world food production. In gen-
eral, temperatures will increase and precipitation patterns will change in midlatitudes
under the influence of climate change, while higher latitudes and tropical regions
will benefit from the change in such conditions. Flat areas in central North and
9  Pest and Disease Management Under Changing Climate 151

South America, southern Europe, southern Australia, and semiarid and subtropical
parts of Asia may experience poor conditions. On the contrary, the regions of
Northern Europe, Canada, Russia, Northeast and Southeast Asia, temperate regions
of America, and mountainous and cold regions will become warmer. Due to changes
in temperature and rainfall, the spread and distribution of cultivation in different
regions will change and the variety of production in these regions will decrease.
Meanwhile, regions with arid and semiarid climates will experience frequent
droughts. It is believed that the temperature of the earth will increase by 0.1  °C
every decade. The stratosphere region will be colder with the increase of carbon
dioxide and the amount of evaporation will increase in the warm climate, which will
cause a global increase in precipitation. At high latitudes and in the Northern
Hemisphere, temperatures and precipitations are expected to increase (Mahlman,
1997). Therefore, the decrease in soil moisture due to the increase in temperature
and precipitation is neutralized. It is expected that as a result of these changes, the
potential of food production will decrease significantly, especially after 2050
(Krishnareddy, 2013). According to the predictions of the Peterson Institute in the
field of agricultural economics, the production of agricultural products in develop-
ing countries will decrease by 10–25% by 2080 (Cline, 2007). With the increase in
population, it is expected that the need for food will double in the middle of the
twenty-first century, while the effects of global warming and water shortage will
reduce production in some areas (Tilman et al., 2002). Therefore, special attention
should be paid to other restrictions on product performance and quality, such as pest
epidemics and emerging pathogens. In this chapter, while examining the impact of
climate change on pests and plant pathogens, the necessary solutions to control pos-
sible consequences are presented.

2 Climate Changes

Global warming increases the level of evaporation and changes the precipitation
pattern. As a result of climate change, in some areas the rainfall will be more and in
some areas the rainfall will be less. Rainfall is expected to be higher in the equato-
rial belt and high latitudes. While for midlatitudes and semiarid and arid regions, a
decrease in precipitation is expected. It is expected that by the end of the twenty-­
first century, the decrease in soil moisture and the frequency of short-term droughts
will double, and long-term droughts will triple, especially in low-rainfall regions
(Füssel, 2009; Solomon et al., 2007).
Climate change is a major threat to ecosystem function and biodiversity and
affects all ecosystem levels (Lepetz et  al., 2009). For example, climate change
directly affects the rate of respiration, photosynthesis, and other biochemical reac-
tions. Absorption of minerals, nutrients, and water, plant canopy exchanges, absorp-
tion of solar energy to produce carbohydrates during the photosynthesis cycle, and
metabolism of carbohydrates in respiration are largely dependent on the climate and
temperature of the environment. Soil biological processes (such as decomposition,
152 Y. Biniaz et al.

mineralization, and stabilization) and soil abiotic processes (including solute trans-
portation, aeration, and ion exchange) are affected by climate change. As the soil
temperature rises, the rate of decomposition of organic materials increases and the
absorption of nutrients by the plant is disturbed. Even if plants have their own mech-
anisms to tolerate adverse conditions, the physiological response of plants in cli-
mate change conditions depends on the limiting factors in the vegetation sites. For
example, an increase in temperature may exacerbate the lack of air vapor pressure.
This process can increase transpiration, especially in dry areas (Amedie, 2013).
Climate change and global warming are largely related to human activities, includ-
ing the emission of greenhouse gases. Destruction of habitats, fragmentation of
native and wild habitats, reduction of biodiversity, and increased risk of pests and
pathogens are considered to be the main impacts of climate change in the agricul-
tural sector (Lepetz et al., 2009).

3 Plant Pests and Diseases

Despite the concerns caused by climate change and population growth, it is surpris-
ing that little research has been done on the impacts of climate change on plant pests
and diseases. Plant diseases are caused by the interaction of sensitive host plant,
pathogen, and environment. Therefore, any change in environmental conditions will
affect the spread of the disease by affecting the host and the pathogen. Changes
related to climate change, such as temperature increase, change in the level and pat-
tern of precipitation, increase in CO2 and destruction of the ozone layer, drought,
etc., may be effective in the emergence and exacerbation of plant pests and diseases
(Chakraborty, 2005; Garrett et al., 2006; Grace et al., 2019). In addition to reducing
crop yields, plant diseases also have a negative effect on the quality and safety of
food. Due to climate change, concerns about pesticides, herbicides, and toxins in
food have increased in Europe. For example, the concentration of fusarium head
blight fungus generally increases with increasing number of rainy days and increas-
ing relative humidity. Therefore, the change in temperature and atmospheric condi-
tions may affect the severity of this disease and the production of nectar (Chakraborty
& Newton, 2011).
Insects are the most diverse category of living organisms on earth. They may
directly or indirectly have a positive or negative impact on human life and natural
ecosystems. Climate and its indicators such as temperature have a strong and direct
effects on the growth, reproduction, and survival of insects (Andrewartha & Birch,
1954; Wang et al., 2022) and climate change is the dominant nonliving factor that
directly affects herbivorous insects (Bale et al., 2002). Considering that the growth
potentials of insect populations are basically controlled by the temperature of the
environment, increasing the temperature may increase or decrease the growth rate
of insects. Generally, damage to the product depends on the optimal temperature
range for different species of insects (Ward & Masters, 2007). The global climate
has experienced many changes in recent years and this process of changes will
9  Pest and Disease Management Under Changing Climate 153

continue in the future. Insects and plant pathogens will probably be affected by
climate change, because their body temperature is affected by the temperature of the
surrounding environment and they are also sensitive to the intensity of rainfall
(Merrill et al., 2008). These impacts may be direct due to the effects of climatic fac-
tors on the physiology and behavior of insects (Thomson et al., 2010), or they may
be indirect due to host plants, competitors, and natural enemies (Torchin et  al.,
2003). Mild winters increase the survival rate of pests and allow them to quickly
restore populations in spring (Bale et al., 2002). In the conditions of climate change,
it is expected that the growing season of plants will be extended, and this is in favor
of multigenerational species so that they can increase the reproduction rate. This
can increase the possibility of damage from invasive species (Ward & Masters,
2007). If average temperatures gradually rise over the long term, then the likelihood
of temperatures cooling enough to kill insects in winter will decrease.
It can be argued that these forces imposed by climate change can incredibly alter
the native pest species as well. The presence of antagonists (effect reducers) and
natural enemies for native pests in domesticated ecosystems is very important
(Torchin et al., 2003). In contrast, invasive pests across geographic ranges may be
more limited by climate change than by biotic interactions (Scherm & Coakley,
2003). Climate change, humidity, temperature, and carbon dioxide levels are aspects
of climate change that most likely change the distribution, establishment, and dam-
age of pests and plant diseases, especially invasive species.

4 Host Plants’ Response to Climate Change

The direct impact of climate change on any plant or plant community may occur in
the absence of pests and pathogens. But it may cause changes in the plant that cause
interaction with pathogens (Garrett et  al., 2006; Juroszek & Tiedemann, 2013).
Changes in the shape of the plant may affect the microclimate and the risk of plant
contamination. In general, the increase in plant density may cause long-term wet-
ting of the leaf surface and contamination with foliar pests and pathogens, but how
nonliving factors interact to infect plants is the key to understanding the impact of
climate change on plants (Mittler, 2006). Abiotic stresses such as temperature and
drought may make the plant more sensitive to pests and pathogens. These factors
may also trigger general defense cycles that lead to increased resistance. Currently,
the results of many researches are in hand, which show that climate change can
cause changes in the stages and rates of pathogen and changes in plant resistance. In
this way, climate change can cause changes in host–pathogen interaction (Coakley
et al., 1999). Climate change can increase or decrease the disease or have no effect
on it (Chakraborty et al., 2000). The metabolic rate of the host plant is one of the
main factors that determine the activity of plant pathogens. Many agricultural prod-
ucts of hot regions show a decrease in yield if the temperature increases, because the
temperature in which they currently live is close to their maximum physiological
temperature (Cerri et al., 2007).
154 Y. Biniaz et al.

High levels of carbon dioxide may lead to changes in plant structure. An increase
in the concentration of carbon dioxide in the earth’s atmosphere can increase the
growth of plants, increase the area of leaves, increase the thickness of leaves,
increase the number of leaves in a plant, and thicken the stems of plants (Lepetz
et al., 2009).
With the increase of photosynthesis in plants, the efficiency of water use in plants
has increased and the damage caused by ozone gas has decreased (Lovisolo et al.,
2003). The effect of high temperature on plants varies throughout the year. In the
cold seasons of the year, heating may reduce the stress in plants, if it increases the
stress in the warmer seasons and months of the year. When high-temperature stress
increases, plant responses may be similar to drought stress and appear with symp-
toms such as wilting, leaf burn, leaf curling, and leaf drop. In addition, in such
conditions, physiological responses such as changes in RNA metabolism, biosyn-
thesis of proteins, enzymes, isozymes, and hormones occur, and for sure, these
changes affect the plant’s sensitivity to pests and pathogens. Climate change may
also be effective in crop production. For example, studies have shown that in the
Philippines, for a degree Celsius increase in temperature, rice yield decreased by
10% (Chakraborty et al., 2000).
The effect of drought stress on the resistance of plants against foliar pathogens
(which have less success in infection in dry conditions) is very complex. For exam-
ple, alfalfa infection with Verticillium albo-atrum fungus shows less symptoms
under drought stress conditions. In some pathogen–host systems, resistance
decreases in drought conditions (Garrett et  al., 2006). Drought can make plants
susceptible to contamination. In addition, flooding can accelerate the spread of
pathogens. In the tropics, tropical species are sensitive to temperature changes due
to their narrow temperature range. These species currently live in conditions very
close to their optimal temperature (Ghini et al., 2011). It seems that due to the indi-
rect effects of drought on plant physiology, the intensity of pests and diseases in
trees increases (Knott et al., 2019).

5 The Impact of Climate Change on Agricultural Products

Climate change is effective in the geographical distribution and growth of plants in


the world. The effect of climate change on plants depends on various factors, such
as the type of plants, their growth cycle (annual or perennial), planting method (self-­
seeding or sowing), and species combinations. It has been estimated that in Australia,
climate change has caused an increase of 30–50% of the crop, which was the result
of an increase in the minimum temperature (Coakley et al., 1999).
It is also claimed that the high amount of carbon dioxide increases photosynthe-
sis and optimum use of water. These two effects (increasing photosynthesis and
optimum use of water) cause an increase in yield in most plants, which is generally
achieved due to an increase in growth. The IPCC has predicted that doubling the
amount of carbon dioxide will increase the yield of plants by an average of 30%.
9  Pest and Disease Management Under Changing Climate 155

According to the IPCC, the level of this increase depends on the latitude of the given
areas. It is possible that the increase in productivity caused by the increase in carbon
dioxide is neutralized by the invasion of insects, plant pathogens, and weeds.
Therefore, it should be noted that in studies of climate change and growth, the issue
of damage by insects, pathogens, and weeds should also be considered (Coakley
et al., 1999).

6 The Direct Effect of Climate Change on Plant–


Pathogen Interaction

Climate change may directly affect various aspects of host plant biology, including
phenology, amount and content of carbohydrates, nitrogen and phenolic com-
pounds, amount of root and stem biomass, number and size of leaves, density of
stomata, conductance of stomata, secretions, and root and affect the composition
and amount of wax on the leaf. An increase in temperature and drought that affects
plant growth may cause changes in the physiology of the host species that alter the
colonization of the host tissue by the pathogen. Abiotic stress may induce general
and active defense pathways in the plant and increase resistance. These stresses can
also increase the plant’s sensitivity to some pathogens. Humidity and temperature
can be effective in the development of pests and diseases by affecting the sensitivity
of the host to pollution. Drought induces plant resistance or plant sensitivity. As a
result, it has an effect on pathogens such as reducing the symptoms of Verticillium
(Garrett et al., 2006) or reducing the resistance of trees against Armillaria in the
forest system (Coakley et al., 1999). Increasing temperature affects the lengthening
of the growing season, plant and pathogen phenology, and genes. For example, Xa7
gene and Pg3 and Pg4 genes in rye are among the genes that react to temperature
changes (Coakley et  al., 1999; Garrett et  al., 2006). Xa7, Pg3, and Pg4 genes are
applied to deal with Xanthomonas oryzae pv. Oryzae and black rust, respectively. It
is predicted that the prevalence and severity of diseases caused by pathogenic fac-
tors will increase under climate change (Wakelin et al., 2018).

7 Response of Vectors and Pathogens to Climate Change

The temperature range of many pathogens for wintering or summering of the patho-
gen and vector is limited due to climatic needs. For example, high winter tempera-
tures (−6 to −10 °C) increase the winter survival of rust-producing fungi (Puccinia
graminis) and as a result increase the disease. Temperature requirements for infec-
tion are different among fungal species. Temperature requirements in wheat rust are
from 2 to 15 °C for yellow rust, −10 to 30 °C for brown rust, and −15 to 35 °C for
black rust (Roelfs, 1992).
156 Y. Biniaz et al.

The arrival of new vector species and the change in their wintering and summer-
ing may have an important impact on the survival of pathogens and their transmis-
sion and reproduction. Temperature is the main climatic factor that has a direct
effect on wintering. Pathogen transmission also takes place in climate change con-
ditions. This issue is more important in the case of viral diseases. Viruses infecting
plants have a close relationship with their host and vector. The risk of local and
regional emergence of vector-borne diseases is related to the climatic needs of the
vector (Malmstrom et al., 2011; Trebicki & Finlay, 2019). In relation to the diseases
that are transmitted by vectors, the increase in temperature often increases the num-
ber of insect generations, and as a result, it will lead to the transmission of viral and
phytoplasma diseases (Dobson, 2009; Robinet et al., 2011). Global warming may
affect the initial infection of the host, the spread of the infection in the host, or the
horizontal transmission of the virus by the vector. Climate change can affect poten-
tial vector geographic range, vector phenology, and overwintering, density, migra-
tion, and vector activity. These factors can alter the degree of stability, rate of
reproduction, virus movement/translocation, and the synergistic relationship
between viruses.

8 Effect of Climate Change on Natural Enemies

Depending on the complexity and species richness, agroecosystems can have good
potential to provide high levels of natural biological control. Therefore, the com-
plexity of the ecosystem can increase the resilience of the ecosystem against the
spread and outbreak of the pests. Although little is known about how climate change
can alter polytrophic levels and competitive interactions, recent studies suggest that
climate change can alter predator–prey relationships due to the high sensitivity of
trophic levels. It is higher due to climatic variability and different optimal tempera-
tures compared to the pests (Voigt et  al., 2003). In this regard, divergence and
branching in the thermal preference of pests and their natural enemies leads to
breaking of temperature or geographic adaptation and increasing the risk of host
pest outbreaks (Hance et al., 2007). Due to changes in the quality and size of the
herbivore host induced by temperature, double negative effects may cause a decrease
in the competence of natural enemies (Thomson et al., 2010). Climatic variability,
in addition to the increase in carbon dioxide, can change and transform the degree
of impact of biocontrol agents by changing the growth, morphology, and reproduc-
tion of the target pest. Populations of natural enemies and host insects may respond
differently to global warming. There is evidence that in warmer climatic conditions,
the degree of impact of many species of natural enemies (Kiritani, 2006) or the
sensitivity of their prey increases (Thomson et al., 2010). In adverse climatic condi-
tions, the high frequency of insect pests may occur to some extent due to the low
activity of parasitoids (Hance et  al., 2007) or the disruption of the relationship
between the parasitoid and the pest, thus reducing the ability to control them. Other
studies have suggested that high temperatures can benefit parasitoids more than
9  Pest and Disease Management Under Changing Climate 157

their hosts (Kiritani, 2006). Insects and parasitoids with low developmental thresh-
old seem to adapt well to existing conditions and their relationships may not be
significantly affected under global warming conditions (Chu & Chao, 2000).
It is clear that the climate and especially the temperature are the most important
factors to limit the distribution, growth rate, number of generations, and abundance
of insect species in a region. In addition, many insect pests can adapt to a wide range
of environments through natural selection and evolution. The degree of effective-
ness of different types of insect pests under the influence of climate change is related
to the degree of change. In other words, the level of changes will have an inverse
correlation with the range of environmental needs. Many insect pests are widely
tolerant and adaptable organisms, and their occurrence in the environment depends
on the presence of a specific host plant. Therefore, they may be less clearly and
distinctly affected by climate change than other species. Climate change may lead
to a change in the occurrence and importance of native insect pests. The abundance
of a number of native insect pests may be in a negative direction (resulting in
reduced harm). The most important reasons for this are frost-free winters, low
humidity, unfavorable weather, an increase in the number of antagonists (effect
reducers), and phenological differences with the host plant. Changes may occur in
host plant–pest interactions. They may occur in certain situations, at increased or
decreased levels of damage.

9 Direct Impact of Climate Change on Pest Host System

Plant pathologists have studied the effect of climatic changes in the environment on
plant diseases for a long time. The classical triangle emphasizes the interaction of
plant hosts, pathogens, and the environment in disease production (Garrett et al.,
2006). Climate change is one of the many ways that the environment can intensify
or reduce the disease in the long term. Therefore, plant diseases can even be consid-
ered as an indicator of climate changes. Of course, there are simpler biomarkers for
this. Long-term information is rare in natural conditions, but if available, it can be a
suitable strategy for assessing the impact of the environment on plant health (Burdon
& Zhan, 2020; Pautasso et al., 2012). Analysis of old samples at the Rothamsted
station in England for a long time (from 1850 onwards) showed that the emission of
SO2 has a significant relationship with the spread of two pathogens, Phaeosphaeria
nodorum and Mycosphaerella graminis (Fitt et al., 2011). Normally, in the condi-
tions of climate change, it can be predicted due to various mechanisms such as
acceleration in pathogen development, shorter incubation time, intensification of
nonliving tensions, and emergence of unfavorable climatic events for the plant. The
reason for this is related to the lack of harmony between the environment and the
climate. It is expected that due to the indirect effect of climate change on the physi-
ological characteristics of the host, dry conditions will be created, which will
increase the frequency of tree pathogens. Also, an increase in temperature causes
more proliferation of vectors, which causes more pathogens to be transmitted.
158 Y. Biniaz et al.

The observations of the warming of the planet cause a change in the host of some
plants to some fungi. More heat in the forests of Canada and America has caused an
outbreak of bark beetles. In early 2000, a severe drought occurred in southern
California. The severe stress caused by mistletoe infection and basidiomycetes rot
made the trees susceptible to attack by bark-eating beetles (Kliejunas, 2010).
Currently, plant pests cause huge damage to agricultural products. An increase in
temperature may increase possible damage to the plants as well. Climate change not
only threatens the health of plants, but may also aggravate disease in some cases
(Fitt et al., 2011). It is predicted that the increase in temperature in northern Germany
will create a good opportunity for oilseed pathogens such as Alternaria brassicae,
Sclerotinia sclerotiorum, and Verticillium longisporum to cause damage. Of course,
this damage will be significant when the average temperature increase in the long
term is considered. Hot climates in forests increase the rate of fire. Although warm-
ing may make biocontrol easier, very little is known about this.

10 Interaction Between Climate Change Factors

The climate change has a direct effect on the increase of pollutants and the level of
carbon dioxide in the climate. This causes the easy entry of foreign invasive species
into the regions. This will endanger the health of the plants. The arrival of new
pathogens has already occurred in many regions, but climate change has made their
establishment and spread easier (Pautasso et al., 2012; Ziska et al., 2018).

11 Disease Prediction Models

Coakley et al. (Power & Mitchell, 2004) have introduced different ways to model
the impacts of climate change on plant diseases. Experimental models are like
regression models in which climate variables are considered as predictors and epi-
demic indicators as response variables. If the mechanism of their relationship is
fully known, it can predict the success of the pathogen in the scope of the studied
conditions. Simulation models are based on theoretical communication and can be
used for prediction in the studied domain.
There are still three main problems for the application of prediction models
caused by climate change in plant diseases. First, the data of the models generally
have high degrees of uncertainty. For example, data on the geographical spread of
the disease is not available. Second, it is difficult to collect data to clearly predict
linear relationships between climate variables and epidemiological responses.
Third, the potential of compatibility between the plant and the pathogen is a com-
plex factor that has been neglected in most models, because climate change happens
slowly and in various ways, it is not possible to study its direct impact. Temperature
variables in the climate can be studied in terms of long-term impact. Access to
9  Pest and Disease Management Under Changing Climate 159

geographic distribution data is still difficult. In addition, uncertain relationship,


relationship threshold between climate variables, epidemiologic responses, and
information preparation are difficult. It should be noted that the compatibility of
pathogen and host has made this challenge more complicated.

12 The Perspective of Pest and Plant Disease Management

Pest and plant disease management solution in climate change needs balance.
Strategies such as delaying cultivation to prevent pests and pathogens are less effec-
tive. Biological control in sustainable agriculture is one of the healthiest methods of
managing pests, pathogens, and weeds. One of the important issues of the applica-
tion of biological control in order to manage plant diseases in natural farm condi-
tions is the danger of the biological control population in the face of environmental
changes and the aggravation of environmental conditions (Stenberg, 2017; Wong
et al., 2002). If the appropriate temperature and humidity are not continuously pro-
vided, the biological control population may be reduced to such an extent that it
does not have a sufficient effective impact. In such a situation, when the conditions
are favorable again, its population may be small compared to the pathogenic popu-
lation (Karlsson Green et al., 2020).
Risk models for the spread of invasive pests and pathogens to new areas are gen-
erally based on climate change indicators such as temperature, rainfall, and humid-
ity. When there is a commercial restriction in the field of diseases such as Indian
leeches in the region, these risk assessment models are of great economic impor-
tance (Garrett et al., 2006). For many invasive pathogens, climate models must be
linked to trade sectors and human networks. The requirement for this is to have
information on the presence of susceptible hosts and the possibility of pathogen
transmission. According to Johnson (Johnson, 1984), stable resistance is the resis-
tance that has long-term and wide application for the pathogen in suitable environ-
mental conditions. If its stable resistance is inherent, climate change has no effect
on it. If the stable resistance is different depending on the conditions, it can be
replaced with a more suitable decision/alternative. If contamination cycles continue
during the growing season due to climate change, stable resistance may be at risk.
The signs of reduced resistance generally include an increase in fertility and an
increase in the number of pest and pathogen generations in the season and favorable
conditions for the development of the disease.
Climate change has two major effects on pesticide use. First, the change in tem-
perature and rainfall affects the residual amount of toxins on the branches. In addi-
tion, continuous rains wash the pesticides off the plant. Second, morphological and
physiological changes in conditions where the concentration of carbon dioxide is
high affect the absorption, release, and metabolism of systemic pesticides. In both
cases, due to the increase in the amount of carbon dioxide, the efficiency of chemi-
cal inhibition decreases (Coakley et al., 1999).
160 Y. Biniaz et al.

Therefore, based on what was said before, it is necessary to find effective ways
to limit pests and diseases and prevent the further development of plant epidemics.
Achieving this goal is possible by knowing the various factors influencing the emer-
gence of pests. Such integrated knowledge enables the integrated management of
pests and plant pathogens and minimizes the effect of selection pressures during this
process (Jones, 2004, 2006). One of the important conclusions about the effects of
climate change on pest and disease management is that climate change leads to an
increase in predictor variables, increasing uncertainty in decision-making. In order
to reduce the negative effects of various human activities, such as the mono-­crop
system, the destruction of genetic diversity, the destruction of native plant cover,
and the excessive use of chemical pesticides, it is necessary to use irrigation and
conservation agriculture in appropriate ways. In addition, it is recommended to
review the consequences of plant quarantine, whose rules have been eased in order
to improve international trade conditions. Quarantine process requires precision,
strictness, and strengthening instead of simplicity, because it plays an important role
in reducing the access to factors influencing the emergence of pests and diseases,
such as the entry of pathogens through plant materials or the transmission of disease
vectors from distant areas.
There are various methods to control the effective factors in the emergence of
pathogens. A combination of actions that is economically viable should be selected.
These measures should be compatible with standard agricultural practices and envi-
ronmentally and socially acceptable (Jones, 2001, 2006). Some of these methods
are completely nonspecific and effective on a wide range of pests and diseases.
Others are used specifically for a specific pathogen. Control measures include dif-
ferent activities that affect the initial inoculum and the rate of epidemic growth.
Agricultural and legal methods are considered nonspecific, while host resistance
and most biological control measures are specific. Chemical control can be specific
or nonspecific depending on the type of pesticide (Jones, 2006; Karlsson Green
et al., 2020). A noteworthy point that increases the importance of nonspecific con-
trol such as agronomic and legal methods is its effect on unknown pathogens or
pathogens on which there is little information (Jones, 2004, 2006; Jones et al., 2004).
In addition to the commonly used methods in managing plant diseases, it is ben-
eficial to use plant breeding to produce plants that meet the needs of the future cli-
mate. Breeding programs can provide plant resistance to diseases, tolerance to
environmental stress conditions, and genetic diversity. Of course, these reform pro-
grams should not ignore the quality and quantity of production. Native microbial
communities play an important role in maintaining plant health. Therefore, promot-
ing these communities has many advantages. With recent advances in technologies
such as metagenomics analysis, knowledge of microbial dynamics in soil and other
environments and the establishment of microbial populations suppressing plant
pathogens will increase. Among the useful microorganisms, there are some of them
that can grow in stressful conditions. Choosing these microorganisms helps farmers
to deal with biotic and abiotic stresses.
9  Pest and Disease Management Under Changing Climate 161

13 Conclusions and Future Perspectives

Climate change is considered as one of the major risks for agriculture and food
security. Therefore, it is necessary to include crops, insect pests, crop pathogens,
and their interactions in future climate change studies. Although many studies have
been conducted in this field, unfortunately, there is still no satisfying answer to the
question of how climate change shapes the relationship between crops, pests, and
pathogens and ultimately affects crop productivity (Wang et al., 2022). Based on
controlled laboratory experiments and models, current knowledge about the impact
of large climatic events on crops, pests, and pathogens is insufficient to describe the
effects of increasing frequency of these events on temporal and spatial scales.
Ignoring important features such as daily maximum and minimum temperatures and
the time elapsed until the next climatic event can lead to misleading models about
the physiological and biochemical responses of the crops, pests, and pathogens to
climate change (Ma et al., 2021). Therefore, future research should investigate the
effects of large climatic events on crops, their pests, and pathogens in more detail.
They can do this in laboratory environments.
Results from current controlled field practices are generally unable to adequately
represent complex real-world conditions (Trębicki et  al., 2017). Also, the results
obtained from local or regional experiments cannot be accurately generalized to
larger scales. Therefore, it is suggested that further large-scale and long-term stud-
ies be conducted to investigate real farm environments in a wide range of latitudinal
and altitudinal gradients.
Interactions between crops, pests, and pathogens are very complex and particu-
larly sensitive to IPM strategies. There are many gaps in existing knowledge regard-
ing the impact of different agricultural practices such as intercropping, changing
planting dates, and the use of pesticides and fungicides on pests and pathogens and
their interaction with crops. Furthermore, regardless of which IPM strategy is
adopted for pest control, these strategies affect beneficial organisms (biological
control agents) and subsequent IPM strategies (Stenberg, 2017). The combined
effect of multiple climatic stressors and IPM strategies on crop production, pests,
and pathogens, and the interaction between them is still unclear. When facing stress-
ful factors, there may be phenomena such as mutual tolerance. In other words,
increasing tolerance to one stressor may increase tolerance to other stressors or vice
versa. Therefore, more research is needed in the field of complex multifactorial
problems to finally realize the full efficiency of IPM. As a result, in circumstances
where climate change and socioeconomic development put maximum pressure on
agricultural products, it is necessary to conduct more studies on the interactions
between crops, pests, pathogens, and agricultural practices under climate change.
This can help farmers and researchers better understand the effects of climate
change on crop production. Only if a complete picture of how climate change affects
crop pests and pathogens is obtained, farmers can take effective measures to control
pests and pathogens, adapt the use of pesticides and fungicides, and reduce the
negative effects of climate change. Although future climate changes are full of
162 Y. Biniaz et al.

uncertainty and difficult to generalize, the present research can provide useful
insights to clarify how insect pests, pathogens, agricultural crops, and their interac-
tions respond to climate change.

References

Amedie, F. A. (2013). Impacts of climate change on plant growth, ecosystem services, biodiver-
sity, and potential adaptation measure [Masters Thesis] Program Study of Biological and
Environmental Science. University of Gothenburg.
Andrewartha, H. G., & Birch, L. C. (1954). The distribution and abundance of animals. University
of Chicago Press.
Bale, J. S., Masters, G. J., Hodkinson, I. D., Awmack, C., Bezemer, T. M., Brown, V. K., Butterfield,
J., Buse, A., Coulson, J. C., Farrar, J., Good, J. E. G., Harrington, R., Hartley, S., Jones, T. H.,
Lindroth, R. L., Press, M. C., Symrnioudis, I., Watt, A. D., & Whittaker, J. B. (2002). Herbivory
in global climate change research: Direct effects of rising temperature on insect herbivores.
Global Change Biology, 8(1), 1–16. https://doi.org/10.1046/j.1365-­2486.2002.00451.x
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting cir-cular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Burdon, J. J., & Zhan, J. (2020). Climate change and disease in plant communities. PLoS Biology,
18(11), e3000949. https://doi.org/10.1371/journal.pbio.3000949
Cerri, C. E. P., Sparovek, G., Bernoux, M., Easterling, W. E., Melillo, J. M., & Cerri, C. C. (2007).
Tropical agriculture and global warming: Impacts and mitigation options. Scientia Agricola,
64(1), 83–99. https://doi.org/10.1590/S0103-­90162007000100013
Chakraborty, S. (2005). Potential impact of climate change on plant–pathogen interactions.
Australasian Plant Pathology, 34(4), 443–448. https://doi.org/10.1071/AP05084
Chakraborty, S., & Newton, A. C. (2011). Climate change, plant diseases and food security: An
overview. Plant Pathology, 60(1), 2–14. https://doi.org/10.1111/j.1365-­3059.2010.02411.x
Chakraborty, S., Tiedemann, A.  V., & Teng, P.  S. (2000). Climate change: Potential impact
on plant diseases. Environmental Pollution, 108(3), 317–326. https://doi.org/10.1016/
s0269-­7491(99)00210-­9
Chu, Y., & Chao, J. (2000). The impact of global change on insects. Journal of Applied Entomology.
National Taiwan University, 341–366.
Cline, W.  R. (2007). Global warming and agriculture: End-of-century estimates by country.
Peterson Institute.
9  Pest and Disease Management Under Changing Climate 163

Coakley, S. M., Scherm, H., & Chakraborty, S. (1999). Climate change and plant disease man-
agement. Annual Review of Phytopathology, 37, 399–426. https://doi.org/10.1146/annurev.
phyto.37.1.399
Dobson, A. (2009). Climate variability, global change, immunity, and the dynamics of infectious
diseases. Ecology, 90(4), 920–927. https://doi.org/10.1890/08-­0736.1
Fitt, B.  D. L., Fraaije, B.  A., Chandramohan, P., & Shaw, M.  W. (2011). Impacts of changing
air composition on severity of arable crop disease epidemics. Plant Pathology, 60(1), 44–53.
https://doi.org/10.1111/j.1365-­3059.2010.02413.x
Füssel, H.-M. (2009). An updated assessment of the risks from climate change based on research
published since the IPCC fourth assessment report. Climatic Change, 97(3–4), 469–482.
https://doi.org/10.1007/s10584-­009-­9648-­5
Garrett, K. A., Dendy, S. P., Frank, E. E., Rouse, M. N., & Travers, S. E. (2006). Climate change
effects on plant disease: Genomes to ecosystems. Annual Review of Phytopathology, 44,
489–509. https://doi.org/10.1146/annurev.phyto.44.070505.143420
Ghini, R., Hamada, E., Pedro Júnior, M.  J., & Gonçalves, R.  R. V. (2011). Incubation period
of Hemileia vastatrix in coffee plants in Brazil simulated under climate change. Summa
Phytopathologica, 37(2), 85–93. https://doi.org/10.1590/S0100-­54052011000200001
Grace, M. A., Achick, T.-F. E., Bonghan, B. E., et al. (2019). An overview of the impact of climate
change on pathogens, pest of crops on sustainable food biosecurity. International Journal of
Ecotoxicology and Ecobiology, 4, 114–119.
Hance, T., van Baaren, J., Vernon, P., & Boivin, G. (2007). Impact of extreme temperatures on par-
asitoids in a climate change perspective. Annual Review of Entomology, 52, 107–126. https://
doi.org/10.1146/annurev.ento.52.110405.091333
Johnson, R. (1984). A critical analysis of durable resistance. Annual Review of Phytopathology,
22(1), 309–330. https://doi.org/10.1146/annurev.py.22.090184.001521
Jones, R.  A. C. (2001). Developing integrated disease management strategies against non-­
persistently aphid-borne viruses: A model programme. Integrated Pest Management Reviews,
6(1), 15–46. https://doi.org/10.1023/A:1020494604184
Jones, R.  A. (2004). Using epidemiological information to develop effective integrated virus
disease management strategies. Virus Research, 100(1), 5–30. https://doi.org/10.1016/j.
virusres.2003.12.011
Jones, R. A. (2006). Control of plant virus diseases. Advances in Virus Research, 67, 205–244.
https://doi.org/10.1016/S0065-­3527(06)67006-­1
Jones, R., Latham, L., & Coutts, B. (2004). Devising integrated disease management tactics
against plant viruses from ‘generic’ information on control measures. Agricultural Sciences,
17, 10–20.
Juroszek, P., & Von Tiedemann, A. (2013). Plant pathogens, insect pests and weeds in a changing
global climate: A review of approaches, challenges, research gaps, key studies and concepts.
Journal of Agricultural Science, 151(2), 163–188. https://doi.org/10.1017/S0021859612000500
Karlsson Green, K., Stenberg, J. A., & Lankinen, Å. (2020). Making sense of integrated Pest man-
agement (IPM) in the light of evolution. Evolutionary Applications, 13(8), 1791–1805. https://
doi.org/10.1111/eva.13067
Kiritani, K. (2006). Predicting impacts of global warming on population dynamics and distri-
bution of arthropods in Japan. Population Ecology, 48(1), 5–12. https://doi.org/10.1007/
s10144-­005-­0225-­0
Kliejunas, J. T. (2010). Review of literature on climate change and forest diseases of western North
America.
Knott, J. A., Desprez, J. M., Oswalt, C. M., & Fei, S. (2019). Shifts in forest composition in the east-
ern United States. Forest Ecology and Management, 433, 176–183. https://doi.org/10.1016/j.
foreco.2018.10.061
Krishnareddy, M. (2013). Impact of climate change on insect vectors and vector-borne plant
viruses and Phytoplasma. In Climate-resilient horticulture: Adaptation and mitigation strate-
gies (pp. 255–277). Springer.
164 Y. Biniaz et al.

Lepetz, V., Massot, M., Schmeller, D. S., & Clobert, J. (2009). Biodiversity monitoring: Some pro-
posals to adequately study species’ responses to climate change. Biodiversity and Conservation,
18(12), 3185–3203. https://doi.org/10.1007/s10531-­009-­9636-­0
Lovisolo, O., Hull, R., & Rösler, O. (2003). Coevolution of viruses with hosts and vectors and
possible paleontology. Advances in Virus Research, 62, 325–379. https://doi.org/10.1016/
s0065-­3527(03)62006-­3
Ma, C.  S., Ma, G., & Pincebourde, S. (2021). Survive a warming climate: Insect responses
to extreme high temperatures. Annual Review of Entomology, 66, 163–184. https://doi.
org/10.1146/annurev-­ento-­041520-­074454
Mahlman, J. D. (1997). Uncertainties in projections of human-caused climate warming. Science,
278(5342), 1416–1417. https://doi.org/10.1126/science.278.5342.1416
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In Climate change:
The social and scientific construct by Suhaib A. Bandh. Springer. Nature Publishing. https://
link.springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Malmstrom, C. M., Melcher, U., & Bosque-Pérez, N. A. (2011). The expanding field of plant virus
ecology: Historical foundations, knowledge gaps, and research directions. Virus Research,
159(2), 84–94. https://doi.org/10.1016/j.virusres.2011.05.010
Merrill, R. M., Gutiérrez, D., Lewis, O. T., Gutiérrez, J., Díez, S. B., & Wilson, R. J. (2008). Combined
effects of climate and biotic interactions on the elevational range of a phytophagous insect.
Journal of Animal Ecology, 77(1), 145–155. https://doi.org/10.1111/j.1365-­2656.2007.01303.x
Mittler, R. (2006). Abiotic stress, the field environment and stress combination. Trends in Plant
Science, 11(1), 15–19. https://doi.org/10.1016/j.tplants.2005.11.002
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Norse, D. (2003). Climate change and agriculture: Physical and human dimensions. Food and
Agriculture Organization and Earthscan Publications.
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Pautasso, M., Döring, T. F., Garbelotto, M., Pellis, L., & Jeger, M. J. (2012). Impacts of climate
change on plant diseases—Opinions and trends. European Journal of Plant Pathology, 133(1),
295–313. https://doi.org/10.1007/s10658-­012-­9936-­1
Power, A.  G., & Mitchell, C.  E. (2004). Pathogen spillover in disease epidemics. American
Naturalist, 164(Suppl. 5), S79–S89. https://doi.org/10.1086/424610
Robinet, C., Van Opstal, N., Baker, R., & Roques, A. (2011). Applying a spread model to iden-
tify the entry points from which the pine wood nematode, the vector of pine wilt disease,
would spread most rapidly across Europe. Biological Invasions, 13(12), 2981–2995. https://
doi.org/10.1007/s10530-­011-­9983-­0
Roelfs, A.  P. (1992). Rust diseases of wheat: Concepts and methods of disease manage-
ment. Cimmyt.
Rosenzweig, C., & Hillel, D. (1998). Climate change and the global harvest: Potential impacts of
the greenhouse effect on agriculture. Oxford University Press.
Scherm, H., & Coakley, S. M. (2003). Plant pathogens in a changing world. Australasian Plant
Pathology, 32(2), 157–165. https://doi.org/10.1071/AP03015
Solomon, S., Qin, D., Manning, M., et  al. (2007). Summary for policymakers. Climate
Change, 1–18.
Stenberg, J. A. (2017). A conceptual framework for integrated pest management. Trends in Plant
Science, 22(9), 759–769. https://doi.org/10.1016/j.tplants.2017.06.010
Thomson, L.  J., Macfadyen, S., & Hoffmann, A.  A. (2010). Predicting the effects of climate
change on natural enemies of agricultural pests. Biological Control, 52(3), 296–306. https://
doi.org/10.1016/j.biocontrol.2009.01.022
9  Pest and Disease Management Under Changing Climate 165

Tilman, D., Cassman, K. G., Matson, P. A., Naylor, R., & Polasky, S. (2002). Agricultural sus-
tainability and intensive production practices. Nature, 418(6898), 671–677. https://doi.
org/10.1038/nature01014
Torchin, M. E., Lafferty, K. D., Dobson, A. P., McKenzie, V. J., & Kuris, A. M. (2003). Introduced
species and their missing parasites. Nature, 421(6923), 628–630. https://doi.org/10.1038/
nature01346
Trebicki, P., & Finlay, K. (2019). Pests and diseases under climate change; its threat to food
security. Wiley.
Trębicki, P., Dáder, B., Vassiliadis, S., & Fereres, A. (2017). Insect–plant–pathogen interactions
as shaped by future climate: Effects on biology, distribution, and implications for agriculture.
Journal of Insect Science, 24(6), 975–989. https://doi.org/10.1111/1744-­7917.12531
Voigt, W., Perner, J., Davis, A. J., Eggers, T., Schumacher, J., Bährmann, R., Fabian, B., Heinrich,
W., Köhler, G., Lichter, D., Marstaller, R., & Sander, F. W. (2003). Trophic levels are differen-
tially sensitive to climate. Ecology, 84(9), 2444–2453. https://doi.org/10.1890/02-­0266
Wakelin, S. A., Gomez-Gallego, M., Jones, E., Smaill, S., Lear, G., & Lambie, S. (2018). Climate
change induced drought impacts on plant diseases in New Zealand. Australasian Plant
Pathology, 47(1), 101–114. https://doi.org/10.1007/s13313-­018-­0541-­4
Wang, B. X., Hof, A. R., & Ma, C. S. (2022). Impacts of climate change on crop production, pests
and pathogens of wheat and rice. Frontiers of Agricultural Science and Engineering, 9(1),
4–18. https://doi.org/10.15302/J-­FASE-­2021432
Ward, N.  L., & Masters, G.  J. (2007). Linking climate change and species invasion: An illus-
tration using insect herbivores. Global Change Biology, 13(8), 1605–1615. https://doi.
org/10.1111/j.1365-­2486.2007.01399.x
Wong, P. T. W., Mead, J. A., & Croft, M. C. (2002). Effect of temperature, moisture, soil type
and Trichoderma species on the. Australasian Plant Pathology, 31(3), 253–257. https://doi.
org/10.1071/AP02020
Ziska, L. H., Bradley, B. A., Wallace, R. D., Bargeron, C., LaForest, J., Choudhury, R., Garrett,
K., & Vega, F. (2018). Climate change, carbon dioxide, and pest biology, managing the future:
Coffee as a case study. Agronomy, 8(8), 152. https://doi.org/10.3390/agronomy8080152
Chapter 10
Climate Change Adaptation Through
Agroforestry: Empirical Evidence
from Indian Eastern Himalayan Foothills

Pritha Datta and Bhagirath Behera

Abstract  Agroforestry is widely recognized as a nature-based solution for building


synergies between climate change adaptation and mitigation in the agricultural sec-
tor. Therefore, understanding the factors influencing farmers’ decision to adopt
agroforestry is critical for achieving sustainability. This study explores farmers’
perceptions of climate change and the functions of agroforestry in coping with cli-
mate risks and, ultimately, employs binary logistic regression to examine household-­
level factors that influence agroforestry adoption in the Indian Eastern Himalayan
foothills. The results show the following: (i) a substantial majority of farmers per-
ceived climate change; (ii) about 47% of farming households adopted agroforestry,
stating trees offer them an alternate income stream in the case of crop failure due to
uncertain rains. Moreover, farmers reported that trees are more resistant to flooding,
less water-intensive, and less prone to pest attacks; and (iii) the likelihood of adopt-
ing agroforestry is significantly influenced by ethnicity, remittances, landholding
size, and the educational attainment of the household head. However, farmers with
secure land tenure and access to extension service agents are less likely to choose
agroforestry over conventional farming of cereals. To enable agroforestry to be
more extensively embraced, the study recommends that farmers’ engagement in
agroforestry should be encouraged through policy interventions and capacity-­
building initiatives.

Keywords  Agroforestry · Adoption · Determinants · Farmers · Himalaya · India

P. Datta (*) · B. Behera


Department of Humanities and Social Sciences, Indian Institute of Technology Kharagpur,
Kharagpur, India
e-mail: bhagirath@hss.iitkgp.ernet.in

© The Author(s), under exclusive license to Springer Nature 167


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_10
168 P. Datta and B. Behera

1 Introduction

Since the industrial revolution, global air temperatures have been rising; however,
most of the warming occurred after 1975 at a rate of almost 0.15–0.20 °C per decade
(NASA, 2022). A million species are in danger of extinction due to biodiversity loss
brought on by the rising temperatures, which also raises the cost of agricultural
production and directly impacts fisheries, cattle, and crop growth (Lehtonen et al.,
2019). The impacts of climate change might even push 100 million people to live
below the poverty line by 2030 (USGLC, 2021). Climate change, according to
Diffenbaugh and Burke (2019), has contributed to at least a 25% increase in the gap
between developed and developing economies since 1960. As a result, climate
change is more than simply a concern for the environment; it also poses a threat to
all humanity because of its disastrous effects, which are already leading to more
instability, especially in developing nations (Malla et al., 2022; Bandh et al., 2021,
2022, 2023; Mushtaq et al., 2020; Parray et al., 2022; Bandh, 2022a, b).
While climate change necessitates multifaceted policy solutions, there is a con-
sensus that the climate system responds to changes in greenhouse gas (GHG) con-
centrations with some delay; so, even if GHG emissions are entirely eliminated, the
effects of climate change will endure (Lanfranchi et al., 2014). Taking the low emis-
sion scenarios into consideration (SSP1-1.9), the “Intergovernmental Panel on
Climate Change” (IPCC) anticipates that temperature will stay well above the recent
decades until at least 2100 (IPCC, 2021). As a result, climate change seems unavoid-
able at this point, and attempting to mitigate or adapt to its effects alone may not be
beneficial on a larger scale. This circumstance calls for developing and implement-
ing such strategies that have the advantages of adaptation and mitigation together.
The agricultural sector is the core for resolving the threats of climate change.
Climate change can result in lower crop yields, while at the same time, land use
changes, forestry, and agriculture account for around 25% of GHG emissions (The
World Bank, 2022). Therefore, this sector is both highly vulnerable to climate
change and a source of greenhouse gas emissions. In this line, a plethora of studies
indicated that agroforestry could be a nature-based solution that can serve both
adaptation and mitigation purposes (Bishaw et al., 2022; Lanfranchi et al., 2014;
Reang et al., 2022). Agroforestry is a term used to refer to “land-use systems and
technologies where woody perennials (trees, shrubs, palms, bamboos, etc.) are
deliberately used on the same land-management units as agricultural crops and/or
animals, in some form of spatial arrangement or temporal sequence” (FAO, 2022).
In other words, agroforestry combines crops and trees in the same land, with or
without animals (Glover et al., 2013). Studies have shown that the trees on the farm
provide some of the vital ecosystem services (e.g., water conservation, improve-
ment in microclimate, enrichment in soil productivity and nutrient, soil conserva-
tion, crop pest, and disease control) and also enhance the socio-economic condition
of the farmers (e.g., produces revenue through the sale of agroforestry items, build-
ing materials, animal feed, traditional medicines, and so on) (Hernández-Morcillo
et al., 2018; Lasco et al., 2014; Quandt, 2020). These services are crucial for increas-
ing farmers’ adaptability and livelihood resilience in the face of climate change
(Quandt, 2020).
10  Climate Change Adaptation Through Agroforestry: Empirical Evidence… 169

Agroforestry also has the potential to be used as a mitigation measure because of


the carbon sequestration in biomass and soil (Lasco et al., 2014). Agroforestry sys-
tems have a much greater capacity to store carbon than conventional agricultural
cropping systems, with an estimated annual carbon storage capacity of
0.29–5.21 Mg ha−1 per year (Dhyani et al., 2021). If 630 million hectares of barren
lands could be converted to agroforestry by 2040, it is estimated that at least 586,000
MgC could be sequestered annually (Abbas et al., 2017). Recognizing its advan-
tages, more than a third of developing nations included agroforestry as a sustainable
way to reduce GHG emissions in their “Nationally Determined Contributions”
(NDCs) following the “Paris Agreement” (Thissen, 2020). Besides, agroforestry
also has a highly positive impact on restricting desertification and land degradation,
as well as enhancing food security (IPCC, 2019). Therefore, both the environment
and the farmers are expected to benefit from this practice (Abbas et al., 2017).
Despite tremendous potentiality, the adoption of agroforestry has remained
highly uneven (Glover et al., 2013). Therefore, understanding the factors influenc-
ing farmers’ decision to adopt agroforestry is critical for achieving sustainability in
this era of climate change. Although past studies (Beyene et al., 2019; Glover et al.,
2013; Jha et al., 2021) have examined several socio-economic determinants of agro-
forestry at the local scale, in praxis, the socio-economic conditions differ from
region to region and therefore the determinants as well. Further empirical research
is required in this situation to elicit more local-level insights. Therefore, in order to
promote the adoption of agroforestry, this study will first assess farmers’ percep-
tions of climate change, then look at how agroforestry can help farmers manage
climate risks, and finally examine household-level factors that affect agroforestry
adoption. The findings of this study could help in designing efficient investment
programs to promote agroforestry and inform policymakers about the factors caus-
ing the disparity in agroforestry adoption.

2 Study Area

The study was carried out in the six villages of the Eastern Himalayan foothills of
the Indian state of West Bengal (Fig. 10.1). This region is located at the base of the
Bhutan Himalayas, and every year floods and flash floods hit this area. Furthermore,
the soil is acidic and porous, and the ability to hold water is low (Datta & Behera,
2022a). From the climatic perspective, this area has a subtropical monsoonal cli-
mate, and the maximum annual rainfall (approximately 78%) occurs during the
monsoon (June–September) (Prokop & Walanus, 2017). This region of West Bengal
experiences more monsoonal rain than the rest of the state, owing to the orographic
effects and the extensive forest areas (Datta & Behera, 2022a).
Recent studies have revealed the impact of climate change on this region, which
primarily manifested as a marked increase in temperature and a decline in monsoon
rainfall, which serves as the key water source for farming (Datta & Behera, 2022b;
Datta & Das, 2019). Additionally, the area is particularly vulnerable to climate
170 P. Datta and B. Behera

Fig. 10.1  Location and physiography of the study area

change due to its socio-economic underdevelopment, which includes a higher per-


centage of scheduled caste and scheduled tribe residents, poverty, lack of education,
and inadequate infrastructural development in the form of industrial, irrigational,
and transportation systems (Datta & Behera, 2022a, b; Jana, 2012).
The six study villages were chosen at random from a list of “climate-induced
vulnerable villages” listed in Alipurduar and Jalpaiguri’s “District Disaster
Management Plan (2020),” viz., Turturi Khanda, Uttar Rangali Baznar, Dalsinghpara,
Ballalguri, Pradhanpara, and Dhumpara (Fig. 10.1). These villages are vulnerable to
flooding, and the inhabitants depend heavily on agriculture for sustenance.

3 Data and Methodology

3.1 Sampling and Primary Data Collection

The primary data used in this study were obtained from 300 farming households in
the selected six study villages (Fig. 10.1). The sample size was calculated in accor-
dance with Cochran (1977), with maximum variability of 50% and a 5.65% allow-
able error at a 95% confidence level. For the interviews, 45 households from each
village were chosen at random. The household surveys were conducted using a
semistructured and pretested schedule from March to July 2021. The head farmer
10  Climate Change Adaptation Through Agroforestry: Empirical Evidence… 171

from the farming households was selected as the interviewee, and in the absence of
the head farmer, the second head was approached. The schedule emphasized data
acquisition on farmers’ socio-economic and personal characteristics, opinions on
climate change, and farming practices. Additionally, it should be mentioned that all
interviews were done in compliance with COVID-19 government regulations.

3.2 Source of Climatic Data

The climatic data (CRU TS v. 4.05) were acquired from the Climate Research Unit
(CRU, https://crudata.uea.ac.uk/). CRU TS is a well-known climate dataset that
covers all terrestrial areas of the world, with the exception of Antarctica only (Harris
et al., 2020). These data are available at a finer resolution of 0.5° * 0.5°, which is
created utilizing the weather station-based observational data from various national
meteorological services and other external sources. Because of its higher resolution
and close correlation with the station records of the India Meteorological Department
(IMD), the CRU dataset has been extensively used to detect meteorological trends
over India (Datta & Behera, 2022b; Robertson et al., 2013).

3.3 Determining the Factors Associated with the Adoption


of Agroforestry

A binary logistic regression (BLR) model was employed to examine household-­


level factors that could influence the adoption of agroforestry. All of the explanatory
variables were chosen based on the uniqueness of the study region, discussions with
agricultural specialists knowledgeable about the local context, and theoretical
underpinnings from prior studies (Table  10.1). More specifications regarding the
BLR model and the formula details are available in the work of Datta and Behera
(2022b). The omnibus test of the model coefficient value in this study was 0.00, and
the chi-square value for the Hosmer-Lemeshow (H-S) goodness-of-fit test was
4.997 with a significance level of 0.758, showing good support for the BLR model
used in this study (Pallant, 2013).

4 Results and Discussion

4.1 Farmers’ Perception of Climate Change and Its


Associated Impacts

Farmers’ perceptions of climate change and its associated impacts are presented in
Fig.  10.2a. A vast majority of the farmers (99.33%) reported an increase in the
annual temperature over the study area during the last three decades (1991–2020),
172 P. Datta and B. Behera

Table 10.1  The choice of explanatory variables


Variables Description Source
Ethnicity Dummy, 1 = tribal, Sood et al. (2008), Ekka (2018)
0 otherwise
Gender of the Dummy, 1 = male, Nkomoki et al. (2018), Beyene et al. (2019), Dhakal
household head 0 otherwise and Rai (2020)
Educational Continuous, years Sood et al. (2008), Islam et al. (2015), Ashraf et al.
attainment of of formal education (2015), Islam et al. (2016), David et al. (2017),
household head Owombo and Idumah (2017), Beyene et al. (2019),
Dhakal and Rai (2020)
Farming Continuous, Owombo and Idumah (2017), Nkomoki et al. (2018)
experience farming experience
in years
Land tenure Dummy, 1 = secure, Sood et al. (2008), David et al. (2017), Nkomoki
0 otherwise et al. (2018), Beyene et al. (2019)
Size of landholding Continuous, total Islam et al. (2015), Ashraf et al. (2015), Islam et al.
landholding size in (2016), Owombo and Idumah (2017), Beyene et al.
hectare (2019), Dhakal and Rai (2020)
Access to irrigation Dummy, 1 = yes, 0 Ashraf et al. (2015), Dhakal and Rai (2020)
otherwise
Non-farm income Dummy, 1 = yes, 0 Sood et al. (2008), Dhakal and Rai (2020), Jha et al.
otherwise (2021)
Remittance Dummy, 1 = yes, 0 Cedamon et al. (2018), Mulyoutami et al. (2020)
otherwise
Access to credit in Dummy, 1 = yes, 0 Ashraf et al. (2015), Owombo and Idumah (2017),
the last 5 years otherwise Nkomoki et al. (2018), Beyene et al. (2019)
Contact with Dummy, 1 = yes, 0 Owombo and Idumah (2017), Dhakal and Rai (2020)
extension agents otherwise

which is consistent with the meteorological trends (Fig. 10.2b). Also, 91.67% of the
farmers perceived a decrease in monsoonal rainfall, and 96.67% expressed that the
monsoon has become more irregular and there is no regularity in its onset and with-
drawal pattern. The analysis of the precipitation record also supports farmers’ per-
ception and shows that precipitation has decreased by about 8.01 mm/year over the
period of 1991–2020 (Fig.  10.2c). The previous studies conducted in this region
also support our findings (Datta & Behera, 2022b; Datta & Das, 2019). Since the
majority of the farming is rainfed in this region, farmers quite easily perceived the
changes in rainfall patterns. Similar to this, Shukla et  al. (2019) also found that
farmers who mainly rely on natural capital for their livelihood are likely to have a
vivid perception of climate change.
As mentioned earlier, the study villages are prone to floods, and 89.67% of the
farmers reported that the intensity of floods is increasing every year. Several studies
are anticipating that climate change will likely exaggerate the disastrous events,
including floods and flash floods in the near future (Mishra et  al., 2019). Sands,
stones, and silt have been continuously encroaching on the farmlands of many
households due to the excessive flow of water from the hilly rivers and streams dur-
ing the monsoon months (June–September). As a result, the farmers experience
10  Climate Change Adaptation Through Agroforestry: Empirical Evidence… 173

Fig. 10.2  Figure showing (a) farmers’ perception of climate change and its impacts, (b) trends in
annual temperature, and (c) trends in monsoon rainfall over the period of 1991–2020

crop loss, land erosion, and decreased soil productivity. Whereas about 74% of the
sample farmers have perceived an increase in pest attacks, especially in their paddy
fields. The high temperature, along with low rainfall during the germination phase
and heavy rainfall at the time of harvesting, causes outbreaks of several crop dis-
eases and pest attacks (Debnath, 2017). A plethora of studies have confirmed that
climate change has increased crop pest and disease occurrences, and with the con-
tinuation of temperature warming, the situation is likely to worsen only (Anik &
Khan, 2012; Sutherst et al., 2011).

4.2 Adoption of Agroforestry and Its Role in Dealing


with the Climate Risks

In the study area, cereal-based farming practices are most prevalent. However, the
practice can become riskier as the effects of climate change linger. In the absence of
technological know-how and poverty, many farmers are constrained to plant just
one or two field crops (paddy and potato), and some flood-affected farmlands are
even left fallow. However, it is promising to see that farmers have started adopting
tree-based agricultural systems, which could be a viable replacement for conven-
tional agriculture. About 47.33% of the sample farmers adopted agroforestry by
planting areca nut (Areca catechu), teak (Tectona grandis), sal (Shorea robusta),
and silk-cotton (Bombax malabaricum) trees in their farmlands (Fig. 10.3). Among
174 P. Datta and B. Behera

Fig. 10.3  Some glimpse of the tree-based farming in the study area: (a) pineapple-areca nut, (b)
teak-sal-taro, (c) paddy-teak-sal-bamboo, and (d) areca nut-black pepper-based agroforestry

these, the areca nut is the tree farmers’ most frequent plant in their agroforestry
systems. It is worth mentioning that not just in the study villages but all over West
Bengal’s sub-Himalayan region, areca nut plantations are currently gaining popu-
larity as a commercial crop.
The sample farmers stated that their planted trees are very low water-intensive,
which has offered adaptation benefits to the decreasing rainfall over the study
region. Also, farmers opined that attacks of pests and diseases are very low in these
trees compared to conventional field crops like paddy. Besides, the woods and fruits
of these agroforestry systems have offered the farmers an alternate income stream
in the case of crop failure due to uncertain rains. Moreover, farmers reported that
trees are more resistant to flooding. In open-ended questions, farmers expressed that
trees provide them with a higher gross return in comparison to paddy. These trees
also do not require the application of chemical fertilizers and pesticides and are not
labor-intensive like paddy; therefore, the cost of cultivation is much lower with a
high return. However, the maturity time for the trees, especially the woody trees, is
much longer, which often causes a long waiting time for payback. Also, sometimes
farmers face price fluctuations of the agroforestry produce and do not get the assured
price. Similar results are also reported by Jaganathan and Nagaraja (2015). However,
farmers never denied that the agroforestry-based cropping system provides
10  Climate Change Adaptation Through Agroforestry: Empirical Evidence… 175

additional income and also acts as income security against instability in the price of
main crop paddy.

4.3 Determinants of the Adoption of Agroforestry

The BLR model is employed to assess the factors that influence the adoption of
agroforestry, and the results are shown in Table 10.2. The analysis shows adoption
of agroforestry is significantly affected by ethnicity, education of the household
head, land tenure, landholding size, access to irrigation, recipient of remittance, and
contact with extension agents. Meanwhile, the gender of the household head, farm-
ing experience, non-farm income, and access to credit in the last 5 years have no
significant impact on the adoption (Table 10.2).
Among the sample farmers, 42% were from tribal communities, and the results
of the BLR model show that ethnicity has a positive impact on the adoption of agro-
forestry. Tribal farmers are 3.24 times more likely to adopt agroforestry than non-
tribal farmers (Table 10.2). The sample villages are dominated by different tribal
communities like Oraon, Munda, Kharia, Karjee, and so on. The traditions and
belief systems of these tribal people are deeply connected with nature. The indige-
nous knowledge of tribal communities associated with nature and trees might have
influenced them to adopt agroforestry in the study area. Furthermore, the literature
on adaptation frequently notes that educated farmers are more conscious of the
impacts of climate change and the need to employ adaptation measures (Ali &
Erenstein, 2017; Datta & Behera, 2022a). In this study, we found that the household
head’s education has a significant positive role in the case of adopting agroforestry.
Our results are similar to the findings of David et al. (2017), who stressed that a
lower level of education or being illiterate limits the adoption rate of agroforestry.

Table 10.2  Descriptive statistics and results of the binary logistic regression
Variables Mean SD Coefficient SE Odds ratio
Ethnicity 0.42 0.49 1.18*** 0.35 3.24
Gender of the household head 0.96 0.20 0.02 0.64 1.02
Educational attainment of household head 2.92 3.73 0.07* 0.04 1.07
Farming experience 30.90 12.27 −0.01 0.01 0.99
Land tenure 0.81 0.39 −2.13*** 0.41 0.12
Size of landholding 0.87 0.99 0.61*** 0.19 1.83
Access to irrigation 0.45 0.50 0.65* 0.37 1.92
Non-farm income 0.73 0.44 −0.12 0.33 0.89
Remittance 0.22 0.41 0.96*** 0.35 2.62
Access to credit in the last 5 years 0.24 0.43 −0.39 0.39 0.68
Contact with extension agents 0.07 0.26 −1.95** 0.82 0.14
Nagelkerke (R2) 0.31
*p < 0.1, **p < 0.05, and ***p < 0.01
176 P. Datta and B. Behera

Additionally, Mensah (2015) emphasized the need for higher education to increase
awareness about climate change, knowledge of its repercussions, and the impor-
tance of adopting sustainable land management practices. However, our results are
dissimilar to Nkomoki et al. (2018) and Dhakal and Rai (2020), who reported no
significant association between education and agroforestry adoption. It should also
be emphasized that farmers’ average year of schooling is only about 3  years
(Table 10.2), which is remarkably low, and thus, significant attention to farmers’
education is required for a higher rate of agroforestry adoption and sustainable land
management.
The adoption of agroforestry is thought to be significantly influenced by the
security of the land tenure. Secure land rights and tenure positively impact the adop-
tion of agroforestry (Owombo & Idumah, 2017). Jha et al. (2021) argue that farm-
ers’ perceptions of somewhat less secure land rights and the fact that agroforestry is
a lengthy venture with a potential waiting period before seeing returns both inhibit
the adoption of agroforestry. However, our research showed that the adoption of
agroforestry is negatively impacted by secure land tenure. About 29% of the sample
farmers lack secure land rights (Table  10.2), and the matter of land rights has
become a persistent and unsolved concern for many people of West Bengal’s hills
and foothills, excluding them from assistance and programs supported by the gov-
ernment (The Statesman, 2020). Therefore, in order to sustain their livelihood
needs, farmers with no secure tenure might have been compelled to choose agrofor-
estry. Furthermore, the successful implementation of land reforms in West Bengal,
which provided greater access to land to landless people through land redistribution
and subsequent livelihood supports in the form of income/employment guarantees,
might have given these farmers a sense of de facto land tenure security over their
land thinking that the state authority may not take their land away in the near future,
which in fact might have encouraged them to invest in agroforestry.
Further, the BLR model shows that the size of landholding has a significant posi-
tive association with the adoption of agroforestry by the farmers, indicating that the
larger the farm size, the higher the likelihood of adoption. Our findings are in agree-
ment with previous studies by Lambert and Ozioma (2011) and Dhakal and Rai
(2020). Farmers with smaller landholdings might not be able to cultivate the usual
cereals and other conventional crops needed to meet household demand because
agroforestry requires a piece of land allocated for at least 7–8 years for a return. On
the other hand, having access to irrigation is one of the most crucial agricultural
endowments because it helps to increase crop productivity. According to our study,
farmers who have access to irrigation are more likely to adopt agroforestry, which
is similar to the findings of Sood and Mitchell (2009), Pingale et al. (2014), and
Dhakal et al. (2015). All these studies indicate that households with irrigation sys-
tems are more likely to adopt agroforestry since growing trees is simply just one
aspect of agroforestry; other irrigation-dependent cereals are also included into the
system. As a result, farmers with access to irrigation might be in a better position to
invest in agroforestry.
Our findings also show that farmers who receive remittances are more likely to
opt for agroforestry than households who do not. Receiving remittance might have
10  Climate Change Adaptation Through Agroforestry: Empirical Evidence… 177

increased farmers’ financial capital and enabled farmers to incorporate agroforestry


into their cropping systems. Cedamon et al. (2017) observed a similar positive asso-
ciation between agroforestry and remittances to our findings. On the other hand,
farmers who interact with extension agents are less likely to favor agroforestry over
traditional farming. Our results are different from those of numerous previous
research, which reported that extension contact increased the level of agroforestry
adoption (e.g., Lambert & Ozima, 2012; Owombo & Idumah, 2017). This could be
because the extension agents in those regions were well trained in agroforestry. In
our study, however, only 7% of farmers had contact with extension agents
(Table 10.2) and claimed that they only sought guidance for upgrading traditional
farming practices.

5 Conclusions and Policy Implication

Agroforestry has a great potential to address the adaptation and mitigation needs in
the agricultural sector. This study is, therefore, undertaken mainly to understand the
determinants of agroforestry adoption in order to facilitate the adoption of agrofor-
estry at a greater rate. The study was conducted at the Eastern Himalayan foothills
of West Bengal, India, which is a climate-vulnerable region. We found that the
adoption rate of agroforestry is still moderate in this area since only 47.33% of the
farmers are currently practicing agroforestry in the study villages. The areca nut-­
based agroforestry systems were the most prominent. Farmers opined that the trees
are low water-intensive and resistant to flood and, at the same time, provide an
alternative income option for the farmers if the main crop, paddy, fails due to rain-
fall uncertainties. Also, the respondents felt that trees are low maintenance as they
require no chemical inputs, and the pest attacks are negligible compared to the con-
ventional field crops. Therefore, agroforestry seems to be an appropriate option for
adapting to the current climate change impacts, i.e., decreased rainfall, irregular
rainfall, floods, and increased crop pest and diseases. Further, the study reveals that
agroforestry adoption is generally driven by a set of socio-economic variables.
Ethnicity, educational attainment of the household head, landholding size, access to
irrigation, and remittance have a significant positive effect on the probability of
adopting agroforestry. Whereas farmers with secure land tenure and contact with
extension service agents are less likely to adopt agroforestry over conventional
farming.
In order to enable agroforestry to be more widely adopted, the study recom-
mends (i) encouraging cooperative-based agroforestry, (ii) enabling an assured mar-
ket to encourage farmers to adopt agroforestry-based cropping systems, (iii)
organizing training and demonstration programs to generate awareness among the
farmers about the scientific rationale behind agroforestry adoption needs, (iv)
capacity-building initiatives to strengthen farmers’ livelihood capitals, and (v)
increasing farmers’ contact with extension agents in order to receive appropriate
information regarding the management of agroforestry. However, our study only
178 P. Datta and B. Behera

explored the drivers of adopting agroforestry and its role in addressing the local
climatic risks. Therefore, future studies might look into the carbon sequestration
potential of the local agroforestry systems, explore climate-suitable tree species for
securing farmers’ livelihoods, and address different environmental concerns.

References

Abbas, F., Hammad, H. M., Fahad, S., Cerdà, A., Rizwan, M., Farhad, W., Ehsan, S., & Bakhat,
H.  F. (2017). Agroforestry: A sustainable environmental practice for carbon sequestration
under the climate change scenarios—A review. Environmental Science and Pollution Research
International, 24(12), 11177–11191. https://doi.org/10.1007/s11356-­017-­8687-­0
Ali, A., & Erenstein, O. (2017). Assessing farmer use of climate change adaptation practices and
impacts on food security and poverty in Pakistan. Climate Risk Management, 16, 183–194.
https://doi.org/10.1016/j.crm.2016.12.001
Anik, S. I., & Khan, M. A. S. A. (2012). Climate change adaptation through local knowledge in the
north eastern region of Bangladesh. Mitigation and Adaptation Strategies for Global Change,
17(8), 879–896. https://doi.org/10.1007/s11027-­011-­9350-­6
Ashraf, J., Pandey, R., de Jong, W., & Nagar, B. (2015). Factors influencing farmers’ decisions to
plant trees on their farmsin Uttar Pradesh, India. Small-scale Forestry, 14, 301–313.
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting cir-cular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Beyene, A. D., Mekonnen, A., Randall, B., & Deribe, R. (2019). Household level determinants
of agroforestry practices adoption in rural Ethiopia. Forests Trees and Livelihoods, 28(3),
194–213. https://doi.org/10.1080/14728028.2019.1620137
Bishaw, B., Soolanayakanahally, R., Karki, U., & Hagan, E. (2022). Agroforestry for sustain-
able production and resilient landscapes. Agroforestry Systems, 96(3), 447–451. https://doi.
org/10.1007/s10457-­022-­00737-­8
Cedamon, E., Nuberg, I., & Shrestha, K. K. (2017). How understanding of rural households’ diver-
sity can inform agroforestry and community forestry programs in Nepal. Australian Forestry,
80(3), 153–160. https://doi.org/10.1080/00049158.2017.1339237
Cedamon, E., Nuberg, I., Pandit, B. H., & Shrestha, K. K. (2018). Adaptation factors and futures
of agroforestry systems in Nepal. Agroforestry Systems, 92(5), 1437–1453. https://doi.
org/10.1007/s10457-­017-­0090-­9
Cochran, W. G. (1977). Sampling techniques (3rd ed). John Wiley & Sons.
10  Climate Change Adaptation Through Agroforestry: Empirical Evidence… 179

Datta, P., & Behera, B. (2022a). Assessment of adaptive capacity and adaptation to climate change
in the farming households of Eastern Himalayan foothills of West Bengal, India. Environmental
Challenges, 7, 100462. https://doi.org/10.1016/j.envc.2022.100462
Datta, P., & Behera, B. (2022b). Do farmers perceive climate change clearly? An analysis of mete-
orological data and farmers’ perceptions in the sub-Himalayan West Bengal, India. Journal of
Water and Climate Change, 13(5), 2188–2204. https://doi.org/10.2166/wcc.2022.058
Datta, P., & Das, S. (2019). Analysis of long-term precipitation changes in West Bengal, India: An
approach to detect monotonic trends influenced by autocorrelations. Dynamics of Atmospheres
and Oceans, 88, 101118. https://doi.org/10.1016/j.dynatmoce.2019.101118
David, M., Bernard, B., & Aringaniza, I. (2017). Determinants of agroforestry adoption as an
adaptation means to drought among smallholder farmers in Nakasongola District, Central
Uganda. African Journal of Agricultural Research, 12(23), 2024–2035. https://doi.org/10.5897/
AJAR2017.12219
Debnath, A. (2017). Sustainable agriculture and food security in the Sundarban region of West
Bengal in the perspective of climate change (Thesis) Submitted for the Degree of Doctor of
Philosophy to Department of Geography. University of Calcutta.
Dhakal, A., & Rai, R. K. (2020). Who adopts agroforestry in a subsistence economy?—Lessons
from the terai of Nepal. Forests, 11(5), 565. https://doi.org/10.3390/f11050565
Dhakal, A., Cockfield, G. & Maraseni, T. N. (2015). Deriving an index of adoption rate and assess-
ing factorsaffecting adoption of an agroforestry-based farming system in Dhanusha District,
Nepal. Agroforestry Systems, 89, 645–661.
Dhyani, S., Murthy, I.  K., Kadaverugu, R., Dasgupta, R., Kumar, M., & Adesh Gadpayle,
K. (2021). Agroforestry to achieve global climate adaptation and mitigation targets: Are South
Asian countries sufficiently prepared? Forests, 12(3), 303. https://doi.org/10.3390/f12030303
Diffenbaugh, N. S., & Burke, M. (2019). Global warming has increased global economic inequal-
ity. Proceedings of the National Academy of Sciences of the United States of America, 116(20),
9808–9813. https://doi.org/10.1073/pnas.1816020116
Ekka, R. (2018). Socio-economic and livelihood analysis of tribals in the Angara Block of
Ranchi District, Jharkhand from agroforestry interventions. Oraon, P.R., Bhat, S.S. and Malik,
M.S. Indian Journal of Agroforestry, 20(1), 23–30.
Food and Agriculture Organization. (2022). Agroforestry definition. Accessed from https://www.
fao.org/. Retrieved 1 Aug 2022.
Glover, E. K., Ahmed, H. B., & Glover, M. K. (2013). Analysis of socio-economic conditions influ-
encing adoption of agroforestry practices. International Journal of Agriculture and Forestry,
3(4), 178–184.
Harris, I., Osborn, T. J., Jones, P., & Lister, D. (2020). version 4. of the CRU TS monthly high-­
resolution gridded multivariate climate dataset. Scientific Data, 7(1), 1–18.
Hernández-Morcillo, M., Burgess, P., Mirck, J., Pantera, A., & Plieninger, T. (2018). Scanning
agroforestry-based solutions for climate change mitigation and adaptation in Europe.
Environmental Science and Policy, 80, 44–52. https://doi.org/10.1016/j.envsci.2017.11.013
IPCC. (2019). Interlinkages between desertification, land degradation, food security and green-
house gas fluxes:synergies, trade-offs and integrated response options. In Climate Change
and Land: IPCC Special Report on Climate Change, Desertification, Land Degradation,
Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial
Ecosystems (pp.  551–672). Cambridge University Press. https://www.ipcc.ch/site/assets/
uploads/sites/4/2019/11/09_Chapter-­6.pdf
IPCC. (2021). The physical science basis. Contribution of Working Group I to the Sixth Assessment
Report of the Intergovernmental Panel on Climate Change. Summary for policymakers. In
V. Masson-Delmotte, P. Zhai, A. Pirani, S. L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen,
L. Goldfarb, M. I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J. B. R. Matthews, T. K. Maycock,
T. Waterfield, O. Yelekçi, R. Yu, & B. Zhou (Eds.), Climate change (pp. 3–32). Cambridge
University Press. https://doi.org/10.1017/9781009157896.001
180 P. Datta and B. Behera

Islam, M. A., Masoodi, T. H., Gangoo, S. A., Sofi, P. A., Bhat, G. M., Wani, A. A., Gatoo, A. A.,
Singh, A., & Malik, A. R. (2015). Perceptions, attitudes and preferences in agroforestry among
rural societies of Kashmir, India. Journal of Applied and Natural Science, 7(2), 976–983.
https://doi.org/10.31018/jans.v7i2.717
Islam, M. A., Sofi, P. A., Bhat, G. M., Wani, A. A., Gatoo, A. A., Singh, A., & Malik, A. R. (2016).
Prediction of agroforestry adoption among farming communities of Kashmir valley, India: A
logistic regression approach. Journal of Applied and Natural Science, 8(4), 2133–2140. https://
doi.org/10.31018/jans.v8i4.1103
Jaganathan, D., & Nagaraja, N. R. (2015). Perception of farmers about arecanut based multispecies
cropping system. Indian Research Journal of Extension Education, 15(2), 49–54.
Jana, A. K. (2012). Backwardness and political articulation of backwardness in the North Bengal
Region of West Bengal. In Rethinking state politics in India (pp. 161–204). Routledge.
Jha, S., Kaechele, H., & Sieber, S. (2021). Factors influencing the adoption of agroforestry by
smallholder farmer households in Tanzania: Case studies from Morogoro and Dodoma. Land
Use Policy, 103, 105308. https://doi.org/10.1016/j.landusepol.2021.105308
Lambert, O., & Ozioma, A.  F. (2011). Adoption of improved agroforestry technologies among
contact farmers in Imo State, Nigeria. Asian Journal of Agriculture and Rural Development,
2(393-2016-23889), 1–9.
Lambert, O., & Ozioma, A.  F. (2012). Adoption of improved agroforestry technologies among
contact farmers inImo State, Nigeria. Asian Journal of Agriculture and Rural Development,
2(1), 1–9.
Lanfranchi. (2014). Economic implications of climate change for agricultural productivity.
M.A.U.R.I.Z.I.O., Giannetto, C.A.R.L.O. and de Pascale, A.N.G.E.L.A. WSEAS Transactions
on Environment and Development, 10, 233–241.
Lasco, R. D., Delfino, R. J. P., Catacutan, D. C., Simelton, E. S., & Wilson, D. M. (2014). Climate
risk adaptation by smallholder farmers: The roles of trees and agroforestry. Current Opinion in
Environmental Sustainability, 6, 83–88. https://doi.org/10.1016/j.cosust.2013.11.013
Lehtonen, A., Salonen, A. O., & Cantell, H. (2019). Climate change education: A new approach for
a world of wicked problems. In Sustainability, human well-being, and the future of education
(pp. 339–374). Palgrave Macmillan.
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In Climate change:
The social and scientific construct by Suhaib A. Bandh. Springer Nature Publishing. https://
link.springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Mensah, E.  J. (2015). Land tenure regimes and land conservation in the African drylands: The
case of northern Ghana. Journal of Land Use Science, 10(2), 129–149. https://doi.org/10.108
0/1747423X.2013.878765
Mishra, A. K., Nagaraju, V., Rafiq, M., & Chandra, S. (2019, June). Evidence of links between
regional climate change and precipitation extremes over India. Weather, 74(6), 218–221.
https://doi.org/10.1002/wea.3259
Mulyoutami, E., Lusiana, B., & van Noordwijk, M. (2020). Gendered migration and agrofor-
estry in Indonesia: Livelihoods, labor, know-how, networks. Land, 9(12), 529. https://doi.
org/10.3390/land9120529
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
National Astronautics and Space Administration. (2022). Global temperatures. Accessed from.
https://earthobservatory.nasa.gov/. Retrieved 1 Aug 2022.
Nkomoki, W., Bavorová, M., & Banout, J. (2018). Adoption of sustainable agricultural practices
and food security threats: Effects of land tenure in Zambia. Land Use Policy, 78, 532–538.
https://doi.org/10.1016/j.landusepol.2018.07.021
Owombo, P. T., & Idumah, F. O. (2017). Determinants of agroforestry technology adoption among
arable crop farmers in Ondo state, Nigeria: An empirical investigation. Agroforestry Systems,
91(5), 919–926. https://doi.org/10.1007/s10457-­016-­9967-­2
10  Climate Change Adaptation Through Agroforestry: Empirical Evidence… 181

Pallant, J. (2013). SPSS survival manual. McGraw-Hill Education.


Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Pingale, B., Bana, O. P. S., Banga, A., Chaturvedi, S., Kaushal, R., Tewari, S., & Neema, S. (2014).
Accounting biomass and carbon dynamics in Populus deltoides plantation under varying den-
sity in terai of central Himalaya. Journal of Tree Sciences, 33, 1–6.
Prokop, P., & Walanus, A. (2017). Impact of the Darjeeling–Bhutan Himalayan front on rainfall
hazard pattern. Natural Hazards, 89(1), 387–404. https://doi.org/10.1007/s11069-­017-­2970-­8
Quandt, A. (2020). Contribution of agroforestry trees for climate change adaptation: Narratives
from smallholder farmers in Isiolo, Kenya. Agroforestry Systems, 94(6), 2125–2136. https://
doi.org/10.1007/s10457-­020-­00535-­0
Reang, D., Nath, A. J., Sileshi, G. W., Hazarika, A., & Das, A. K. (2022). Post-fire restoration of
land under shifting cultivation: A case study of pineapple agroforestry in the Sub-Himalayan
region. Journal of Environmental Management, 305, 114372. https://doi.org/10.1016/j.
jenvman.2021.114372
Robertson, A. W., Bell, M., Cousin, R., Curtis, A., & Li, S. (2013). Online tools for assessing the
climatology and predictability of rainfall and temperature in the Indo-Gangetic Plains based
on observed datasets and seasonal forecast models [CCAFS working paper].
Shukla, R., Agarwal, A., Sachdeva, K., Kurths, J., & Joshi, P. K. (2019). Climate change percep-
tion: An analysis of climate change and risk perceptions among farmer types of Indian Western
Himalayas. Climatic Change, 152(1), 103–119. https://doi.org/10.1007/s10584-­018-­2314-­z
Sood, K. K., & Mitchell, C. P. (2009). Identifying important biophysical and social determinants
of on-farm tree growing in subsistence-based traditional agroforestry systems. Agroforestry
Systems, 75(2), 175–187. https://doi.org/10.1007/s10457-­008-­9180-­z
Sood, K. K., Najiar, C., Singh, K. A., Handique, P., Singh, B., & Rethy, P. (2008). Association
between socio-economic parameters and agroforestry uptake: Evidences from eastern
Himalaya. Indian Journal of Forestry, 31(4), 559–564.
Sutherst, R. W., Constable, F., Finlay, K. J., Harrington, R., Luck, J., & Zalucki, M. P. (2011).
Adapting to crop pest and pathogen risks under a changing climate. WIREs Climate Change,
2(2), 220–237. https://doi.org/10.1002/wcc.102
The Statesman. (2020). Bista urges CM to provide land rights to people. https://www.thestates-
man.com. Retrieved 6 July 2022.
The World Bank. (2022). Agriculture and food. Accessed from. https://www.worldbank.org/.
Retrieved 1 Aug 2022.
Thissen, W. (2020). Why agroforestry is a promising climate change solution. Accessed from.
https://www.renature.co/. Retrieved 31 July 2022.
USGLC. (2021). Climate change and the developing world: A disproportionate impact. Accessed
from. https://www.usglc.org/. Retrieved 1 Aug 2022.
Chapter 11
Policy Framework to Introduce
Climate-­Smart Agriculture

Fatemeh Fathi , Naser Valizadeh , Samira Esfandyari Bayat ,


and Khadijeh Bazrafkan

Abstract  Climate-smart agriculture (CSA) helps to increase the resilience of agri-


cultural communities to the impacts of climate change and reduce greenhouse gas
emissions. Moving to the CSA requires very strong policy commitment and deter-
mination. However, this political determination is not possible without the coher-
ence and coordination between the various sectors and stakeholders that are
struggling with climate change. In other words, the coherence of cross-sectoral poli-
cies and programs is the cornerstone of CSA development. In this regard, the devel-
opment of a policy framework to introduce CSA was determined as the main goal
or the main fundamentals of any program for introducing CSA. This framework can
facilitate the achievement of more efficient, adaptable, and flexible agriculture in
the face of climate change. Furthermore, efficient and effective operation of the
CSA elements including CSA techniques, practices, methods, innovations, interac-
tions, arrangements, etc. depends on the existence of these four fundamental
requirements. The use of this framework can provide a basis for easier introduction
of CSA in agricultural societies, reduction of greenhouse gas emissions, and mitiga-
tion of the effects of climate change.

Keywords  Climate-smart agriculture · Policy framework · Multistakeholder


systems · Decision-making tool

F. Fathi
Department of Agricultural Economics, School of Agriculture, Shiraz University, Shiraz, Iran
N. Valizadeh (*) · K. Bazrafkan
Department of Agricultural Extension and Education, School of Agriculture, Shiraz
University, Shiraz, Iran
e-mail: n.valizadeh@shirazu.ac.ir
S. Esfandyari Bayat
Department of Agricultural Extension and Education, College of Agriculture, Tarbiat
Modares University (TMU), Tehran, Iran

© The Author(s), under exclusive license to Springer Nature 183


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_11
184 F. Fathi et al.

1 Introduction

Climate change affects the agricultural sector due to high temperatures, more vari-
able rainfall, and climatic events such as droughts, water shortages, heat waves,
rising sea levels, the spread of pests and diseases, and extinction of biodiversity
(Azadi et al., 2021). The agricultural sector has a very important and direct role in
the emission of greenhouse gases and it is estimated that about 14% of the emission
of greenhouse gases is related to this sector. In addition, the agricultural sector con-
tributes to intensifying the emission of greenhouse gases by changing land use and
deforestation (Seebauer, 2014). In this regard, the adaptation of the system and
activities of the agricultural sector to climate change is one of the undeniable neces-
sities to reduce the impacts of climate change. It is necessary to take practical steps
in the agricultural sector to mitigate the effects of climate change and create more
adaptation to it (Woolf et al., 2018). Climate-smart agriculture (CSA) is one of the
promising alternatives proposed to deal with these crises in the agricultural sector
(Mwongera et al., 2017; Malla et al., 2022; Bandh, 2022a, b; Bandh et al., 2021,
2022, 2023; Mushtaq et al., 2020; Parray et al., 2022).

2 The Evolution of Climate Policies and the Emergence


of an Agricultural System Called CSA

In general, the historical evolution of CSA can be studied in different phases. One
of the most fundamental works in the field of CSA evolution has been done by
Gubta (2010). This researcher sees the evolution of CSA through the lens of inter-
national climate change policies and divides it into five main periods/phases: prob-
lem framing, leadership articulation, conditional leadership, leadership competition,
and leadership during recession (focusing on developing countries). The first phase
in the evolution of CSA is related to the years before 1990. This period is the phase
of framing the problem. In this period, very special events occur in the field of cli-
mate change policies. For example, in 1979 the World Climate Conference was
held. In addition, a strategic study center for climate change studies was established,
which played a key role in explaining global macro policies for climate change in
the following years. This strategic center was the International Panel on Climate
Change (IPCC). It was established in 1988 by the World Meteorological Organization
(WMO) and the United Nations Environment Program (UNEP) and later endorsed
by the United Nations General Assembly. One of the main challenges of the climate
change policy in this period was the lack of a logical connection between the goals
of sustainable development and climate change. However, the Brantland Conference
in 1987 and the publication of its statement covered this perspective gap. Basically,
until then, climate change was not considered as one of the development problems
of countries. In other words, one of the key results of the Brantland Conference
statement was that the efforts of countries to reduce the effects of climate change
11  Policy Framework to Introduce Climate-Smart Agriculture 185

through the reduction of greenhouse gas emissions is one of the necessities of sus-
tainability in development. Therefore, it can be mentioned that the most important
goal of climate change policies in this period was to increase understanding about
the need for collective actions to reduce the effects of climate change by reducing
greenhouse gas emissions (Lipper & Zilberman, 2018).
In the second period, which corresponds to the years 1991 to 1996 (leadership
articulation), several measures were taken in the direction of greater coherence and
coordination in climate change policies among different countries of the world. In
1992, the Rio Declaration on Environment and Development, Agenda 21, the
Biodiversity Convention, and the United Nations Framework on Climate Change
were accepted by many countries. One of the prominent features of this period was
the cooperation of countries to define and quickly implement regulations and proto-
cols related to reducing the effects of climate change and greenhouse gas emissions.
Considering that the contribution of different countries to negative impacts of cli-
mate change was very different, developed countries promised leadership during
this period. This leadership focused more on financial aid to developing countries,
actions that developed countries should take at home, and creating space for devel-
oping countries’ greenhouse gas emissions to grow. The adoption of laws related to
climate change in many countries (including developing countries) led to the accep-
tance of the leadership paradigm of these countries in the leadership articulation
period. Therefore, these countries were recognized as leader countries rather than
polluters. Although the elements of the leadership paradigm were expressed in this
period, their content was more rhetorical than substantive. The third phase of evolu-
tion, which includes the period from 1996 to 2021, is known as the conditional
leadership period. One of the key elements in this period was the change in the
nature of the leadership discourse that was raised in the previous period. In the
explanation of the elements of leadership in the second period, it was stated that
developed countries should help developing countries in reducing greenhouse gases
and take practical measures in the field of reducing/modifying industrial activities
that lead to the emission of greenhouse gases. But this element was violated by
some developed countries such as the United States of America. In this period, the
leadership paradigm became conditional leadership. In other words, unlike the pre-
vious period, developed countries have less responsibility for their activities that
lead to the emission of greenhouse gases. In return, they pledged to compensate for
this reduced commitment by investing in other countries. It should be noted that
during this period, political and economic issues were analyzed more realistically.
Also, many challenges were raised in the context of exempting some developing
countries from meaningful measures in the field of greenhouse gas reduction (Gubta,
2009). The fourth period is from 2002 to 2007. The discourse of leadership was
seriously challenged during this period. Because the United States of America did
not adhere to the Kyoto Protocol, many countries had little desire to fulfill their
obligations to reduce greenhouse gases and climate measures (Lipper & Zilberman,
2018). This led to increased competition among countries in the field of climate
change political strategies (Lipper & Zilberman, 2018; Tan, 2010). Finally, the fifth
period is related to 2008 onwards, which is the period of recession or economic
186 F. Fathi et al.

crisis. One of the key points in this period is that the economic recession has created
other priorities. This factor has caused a decrease in attention to measures to deal
with climate change (Gubta, 2010). Nevertheless, most of the attention was focused
on the adaptation of the agricultural sector to climate change and mitigation from
agriculture. As one of the useful solutions for agricultural adaptation to climate
change and mitigation, CSA has been introduced (Lipper & Zilberman, 2018). CSA
can not only play a key role in reducing the effects of climate change, but can also
help reduce poverty and environmental health (Chandra et al., 2018; FAO, 2010).

3 Definition of CSA

So far, many definitions for CSA have been provided by organizations of climate
change, agricultural and environmental researchers, and international scientific
institutions around the world. For example, in a more general perspective, Adesipo
et al. (2020) define CSA as a form of transformed agriculture that tries to increase
productivity in production systems and food security by using pillars of climate
change and smart and advanced technology. The pillars of climate change in the
definition of these researchers refer to adaptation, resilience, and mitigation. In the
end, they state that this farming approach can increase profit and reduce the vulner-
ability of various sectors, including farmers, by reducing greenhouse gas emissions.
CSA is known as an advanced agricultural system that has three major capabilities,
including a significant increase in crop yield, an increase in the ability to deal with
severe climatic events, and an increase in the effectiveness of coping strategies with
climate change. In fact, these agricultural systems are basically developed to opti-
mize the use of inputs and postharvest agricultural management activities (Chandra
et al., 2018). This agricultural system is different from the usual and old agricultural
systems. However, in this agricultural system, the coordination of the activities of
different sectors and people such as farmers, researchers, private sector, governmen-
tal sector, civil societies, managers, practitioners, policy-makers, and change agents
is at a maximum level. This is considered a basic necessity to create adaptation
capacity, adapt agricultural production systems to climate changes, and increase the
efficiency of using resources and inputs (Azadi et al., 2021; Rosenstock et al., 2016).
Engel and Muller (2016) claim that CSA is a sustainable approach that can
increase agricultural income and productivity by adopting new adaptation technolo-
gies. In addition, they state that this helps to improve the resilience of the agricul-
tural sector against climate change and reduce the emission of greenhouse gases
from the agricultural sector. Adesipo et al. (2020) state that CSA is an approach to
agriculture in which the agricultural system of farmers is prepared to achieve goals
such as sustainability, reducing greenhouse gas emissions, reducing vulnerability,
increasing profit, food security, and increasing productivity. Of course, they state
that this requires capacity building, knowledge, and new technology. In another
definition, Campbell et al. (2014) define CSA as strategies that prepare agricultural
systems under climate change conditions to increase food production and achieve
11  Policy Framework to Introduce Climate-Smart Agriculture 187

food security for the growing world population. CSA provides enabling tools
through which new practices and technologies can be evaluated based on their out-
puts for national development and food security goals under climate change. Also,
it should be emphasized that this approach to agriculture combines sustainable agri-
cultural development, new knowledge, and top-down collaborative approaches
(Nagothu & Kolberg, 2016). CSA considers sustainable intensification in the agri-
cultural sector as one of the main foundations to support proper productivity and
income. Of course, it should be noted that this work must be accompanied by mea-
sures to protect agricultural lands. CSA emphasizes on the application of low-­
income agricultural activities such as conservation agriculture, agroecology,
environmental services, irrigation in small-scale agricultural lands, aquaculture and
agroforestry systems, water conservation, soil conservation, nutrient management,
animal husbandry, integrated management, and reduced tillage for increasing pro-
ductivity and adapting to climate change (Chandra et al., 2018). In general, CSA
can be defined as an integrated approach to develop technical, policy, and invest-
ment conditions to achieve the goals of sustainable agricultural development and
food security under climate change (Bhattacharyya et al., 2020; Food and Agriculture
Organization, 2013).

4 Opportunities in Moving Towards CSA

Investing in CSA can provide many opportunities for different countries to more
easily develop and achieve food security for their growing population. Researchers
have pointed out different opportunities for CSA in their studies. Bhattacharyya
et al. (2020) state that CSA has seven major advantages. First of all, there is a lot of
financial support from CSA in the world. For example, the United Nations is one of
the organizations that supports the development of this agricultural approach very
well. Second, currently more emphasis is placed on government investments. Third,
after the declarations of the Rio, Paris, and IPCC conferences, more emphasis has
been placed on the dilemmas of CSA. Fourth, the new form of the Green Climate
Fund has created an opportunity to support CSA adaptation strategies. Fifth, the
world’s environmental facilities are moving towards creating synergy in the field of
climate change mitigation and adaptation. Sixth, different countries around the
world are easing the rules for the development of CSA. Seventh, carbon trading can
be a good financial deterrent for CSA. Hellin and Fisher (2019) state that CSA and
activities related to this agricultural system can lead to increased adaptive capacity.
In other words, increasing the capacity of adaptation is one of the opportunities that
can be taken advantage of by shifting towards this agricultural system. They also
identify achieving food security goals and climate change mitigation as two other
CSA opportunities. According to Ali and Erenstein (2017), reducing the negative
effects of climate change on the agricultural sector is one of the opportunities cre-
ated by implementing CSA. Lipper et al. (2014) state that increasing coordination
and convergence between different stakeholders, increasing the efficiency of
188 F. Fathi et al.

resource use, and adapting agricultural production systems are among the opportu-
nities of CSA for the agricultural sector.
Increasing productivity, adaptability, and reducing greenhouse gases and carbon
footprints are other opportunities for CSA development. Chandra et al. (2018) men-
tion that CSA has three opportunities including increasing the production of agri-
cultural products, increasing the ability to deal with climatic events, and contributing
to climate change solutions. These researchers also consider the optimization of
input consumption and more effective management of postharvest activities as
potential opportunities for CSA. In general, the most important CSA opportunities
are presented in Table 11.1.

5 Challenges of Moving Towards/Establishing CSA

In contrast to the opportunities created by CSA, this agricultural system also faces
challenges. Bhattacharyya et al. (2020) have raised seven challenges for CSA. First,
in this agricultural system, little investment is made on agriculture itself. In other
words, most of the funds are spent on buying and building new tools and measures.
Second, most investments are made by farmers, foresters, and other vulnerable
groups. This shows that investors are often small or from the private sector. Third,
due to the fact that most of its emphasis is on the sustainability of agricultural activi-
ties and the achievement of food security, and little attention is paid to the profits of
investors, private investors are very interested in it. Fourth, financial support is
focused on mitigation in the energy and industrial sectors. Fifth, most investments
are generally focused on mitigation and adaptation. Sixth, the voluntary carbon
market does not exist due to high transaction costs and lack of proper mechanisms
for quantifying stored carbon. Seventh, in many developing countries, special insti-
tutional facilities for CSA are not yet provided.
Harvey and colleagues (2014) claim that the size of agricultural land has always
been one of the main challenges of adopting new practices and technologies.
Farmers who have large agricultural land are more likely to adopt CSA. However,
they can use a part of their agricultural land to test the efficiency of new techniques
and technologies and use the other part for their traditional activities (Josephson
et al., 2014). This is despite the fact that small-scale farmers are less willing to use
CSA activities. This problem is related to their uncertainty about the expected
results and profit from this agricultural system (Abera et  al., 2020). Land tenure
systems are another challenge of CSA development, especially among small-scale
farmers. In some countries of the world, farmers have the right to use land to develop
agricultural activities, but they cannot own it. This leads to a decrease in farmers’
motivation to invest in innovative activities such as CSA (Nigussie et  al., 2015;
Zeng et al., 2018).
Lack of knowledge in the field of CSA is another point that can be a challenge
for the development of CSA. CSA is an agricultural approach that is based on inten-
sive knowledge. In other words, all the elements of agriculture in this approach,
11  Policy Framework to Introduce Climate-Smart Agriculture 189

Table 11.1  Opportunities of developing CSA


Nom. Opportunity Source
1 Reducing the negative effects of Ali and Erenstein (2017)
climate change on the agricultural
sector
2 Increasing productivity Chandra et al. (2018), Nagothu and Kolberg
(2016), Engel and Muller (2016), Adesipo et al.
(2020), and Campbell et al. (2014)

3 Shifting attention to CSA dilemmas Bhattacharyya et al. (2020)


4 Contributing to sustainability Adesipo et al. (2020)
5 Synergy in the field of climate Bhattacharyya et al. (2020)
change mitigation and adaptation
6 Achieving food security goals Hellin and Fsher (2019), Azadi et al. (2021),
Nagothu and Kolberg (2016), Adesipo et al.
(2020), and Campbell et al. (2014)
7 Using carbon trading as a good Bhattacharyya et al. (2020)
financial disincentive for CSA
8 Increased adaptive capacity Hellin and Fsher (2019), Lipper et al. (2014), and
Chandra et al. (2018)
9 Facilitate rules for the development Bhattacharyya et al. (2020)
of CSA by countries
10 Mitigation of climate change Hellin and Fsher (2019) and Engel and Muller
(2016)
11 Being financially supported by a lot Bhattacharyya et al. (2020)
of organizations
12 Increasing coordination and Lipper et al. (2014)
convergence between different
stakeholders
13 Increasing the efficiency of using Lipper et al. (2014)
resources
14 More emphasis on government Bhattacharyya et al. (2020)
investments
15 Reducing greenhouse gases and Engel and Muller (2016), and Adesipo et al.
carbon footprint (2020)
16 Contributing to climate change Chandra et al. (2018)
solutions
17 Optimizing the consumption of Chandra et al. (2018)
institutions
18 More effective management of Chandra et al. (2018)
postharvest activities
19 Increasing income Engel and Muller (2016) and Adesipo et al.
(2020)
20 The new form of the Green Climate Bhattacharyya et al. (2020)
Fund
21 Reducing vulnerability Adesipo et al. (2020)
22 Providing tools for evaluating the Campbell et al. (2014)
technologies and practices
(continued)
190 F. Fathi et al.

Table 11.1 (continued)
Nom. Opportunity Source
23 Combining sustainable agricultural Nagothu and Kolberg (2016)
development, new knowledge, and
top-down collaborative approaches
24 Emphasis on low-income agricultural Chandra et al. (2018)
activities

such as the level of rains, the state of climatic events, farm management, required
technologies, air and soil moisture levels, soil nutrient balance, etc., are based on
knowledge (Deressa et  al., 2010; Mohammed, 2016; Zerga & Gebeyehu, 2016).
This is despite the fact that in many cases farmers do not have enough knowledge
and information in this field, and this leads to nonacceptance of CSA measures and
technologies (Ogato, 2014). It should be noted that the transfer of this knowledge is
usually done by knowledge transfer systems, agricultural extension, and agricul-
tural advisory services. In other words, it can be said that the lack of a strong and
supportive knowledge transfer system for farmers can be another obstacle to the
development of CSA (Zerssa et al., 2021). Issues related to the culture of agricul-
tural communities can be considered as another challenge to the development of
CSA. For example, gender is one of the things that can affect the development of
CSA. However, the development of technologies in rural communities may be gen-
dered. In other words, many technologies are generally available to men, and women
have little access to technologies, knowledge, and training related to CSA (Josephson
et al., 2014; Lima, 2016).
Lack of financial return in the short term is another challenge of CSA. This prob-
lem is exacerbated when the target population of CSA is farmers who are smallhold-
ers and generally financially weak. This makes many of them look for agricultural
activities that have immediate benefits for them. This is because CSA practices
generally have good financial returns for farmers in the long run (Aweke, 2017;
Zerssa et al., 2021). In such a situation, the livelihood of farmers and their families
will face problems (Aweke, 2017). In such a situation, providing financial aid from
the government or the private sector can have a significant effect on reducing the
vulnerability of farmers’ households (Branca et  al., 2012). But, unfortunately, in
many underdeveloped and developing countries, farmers are generally deprived of
such financial assistance. In other words, the government does not have the ability
to provide sufficient financial aid for them (Deresse & Zerihun, 2018). Therefore,
instead of using CSA, farmers prefer to use their traditional and previous farming
systems, which are less expensive for them. In general, the most important chal-
lenges of CSA are presented in Table 11.2.
11  Policy Framework to Introduce Climate-Smart Agriculture 191

Table 11.2  Challenges of developing CSA


Nom. Challenge Source
1 Low investment in agriculture itself Bhattacharyya et al. (2020)
2 Lack of knowledge in the field of CSA Deressa et al. (2010), Mohammed
(2016), Zerga and Gebeyehu (2016),
and Ogato (2014)
3 Emphasizing on the sustainability of agricultural Bhattacharyya et al. (2020)
activities and achieving food security and paying
little attention to the profits of investors
4 Lack of financial return in the short term Aweke (2017), Zerssa et al. (2021),
and Branca et al. (2012)
5 Focusing investments on mitigation and adaptation Bhattacharyya et al. (2020)
6 Absence of voluntary carbon market Bhattacharyya et al. (2020)
7 Absence of special institutional facilities for CSA Bhattacharyya et al. (2020)
8 The small size of agricultural lands Harvey et al. (2014)
9 Land tenure systems Nigussie et al. (2015) and Zeng
et al. (2018)
10 Investment by vulnerable groups Bhattacharyya et al. (2020)
11 Lack of a strong and supportive knowledge Zerssa et al. (2021)
transfer system for farmers
12 Cultural issues Lima (2016) and Josephson et al.
(2014)
13 Focusing financial support on mitigation in the Bhattacharyya et al. (2020)
energy and industrial sectors
14 Failure in providing sufficient financial aid from Deresse and Zerihun (2018)
the government or the private sector

6 CSA Technologies/Activities

Very diverse methods for CSA have been introduced in different studies. These
practices generally include various on-farm practices that contribute to the sustain-
ability of agricultural activities in the face of climate change. Among these prac-
tices, we can mention agronomy, agroforestry, animal husbandry, forestry, land use,
livestock grazing, water and soil management, and bioenergy (Bryan et al., 2013;
Thorn et al., 2016). Basically, CSA practices should be able to help to achieve the
goals or three dimensions of CSA. These dimensions include increasing production
in a sustainable manner, supporting farmers’ adaptation to climate change and its
effects, and reducing the level of greenhouse gas emissions (Harvey et al., 2014;
Lipper et al., 2014). However, it should not be forgotten that in some cases, many
agricultural practices may be included as CSA supporting practices. This is despite
the fact that many of them do not contribute much to the realization of CSA goals in
practice. In such cases, old agricultural activities are rebranded as CSA supporting
agricultural activities (Chandra et  al., 2018; Ewbank, 2015). Partey et  al. (2018)
classify CSA practices in five categories, which include climate information ser-
vices, conservation agriculture, drip irrigation, agroforestry, planting pits, and
192 F. Fathi et al.

Table 11.3  CSA practices


Nom. Practice Source
1 Conservation agriculture Partey et al. (2018)
2 Agroforestry Vaast et al. (2016)
3 The use of organic manure Ashraf et al. (2021)
4 Crop rotation Mango et al. (2018)
5 Crop diversification Subedi et al. (2019)
6 Mulching Abegunde et al. (2019)
7 Use of wetland FAO (2014)
8 Use of drought- and heat-tolerant crops Ojoko et al. (2017)
9 Use of cover crops Ojoko et al. (2017)
10 Soil conservation techniques Ojoko et al. (2017)
11 Integrated crop-livestock management Abegunde et al. (2019)
12 Improved grazing García de Jalón et al. (2017)
13 Efficient manure management García de Jalón et al. (2017)
14 Diet improvement for animals García de Jalón et al. (2017)

erosion control techniques. These researchers then addressed the issue of which of
these three dimensions of CSA are more related to each of these activities. In gen-
eral, Abegunde et al. (2019) summarize CSA activities in Table 11.3:

7 Positioning Appropriate Policy in Moving Towards/


Establishing CSA

Moving towards CSA requires the use of different methods and techniques that
provide farmers with the possibility to deal with climate change and its effects.
However, these methods and techniques alone cannot provide the basis for the suc-
cess of CSA. In other words, to achieve its goals, this agricultural approach requires
strong political commitments and greater coherence, coordination, and integration
among different sectors related to climate change, agricultural development, and
food security. Therefore, the expansion of CSA to increase the resilience of agricul-
tural communities to the impacts of climate change and reduce greenhouse gas
emissions is highly dependent on the coherence of national policies and intersec-
toral planning. Identifying converging points and strengthening synergies between
different policy goals, dealing with trade-offs, and taking compensatory measures
are among the different goals of the policy system in the field of CSA. Understanding
local barriers and gender discrimination is essential to adopting climate-smart prac-
tices. In addition, incentive mechanisms that can facilitate their adoption can help
policy-makers to design new climate-sensitive policies if necessary (FAO, 2019).
11  Policy Framework to Introduce Climate-Smart Agriculture 193

8 Policies Guiding CSA at the Global Level

In general, it can be mentioned that there are three milestones in the history of cli-
mate change that give direction to CSA policies. In 2015, two major international
agreements were reached that affect CSA policies, strategies, and actions from the
global to the local levels. First, the 2030 Agenda for Sustainable Development and
the Sustainable Development Goals (SDGs) were adopted in September 2015.
Second, the Paris Agreement on Climate Change was reached in December 2015
and ratified in November 2016. The third global framework that guides national
sustainable development activities in the field of CSA is the Sendai Framework for
Disaster Risk Reduction, which was adopted at the Third World Conference on
Disaster Risk Reduction in Sendai, Japan, in March 2015. There are other interna-
tional agreements that may not have an overt climate focus, but may nevertheless
influence national climate policy objectives (FAO, 2019).

9 The Elements of a Policy System for CSA


at the National Level

CSA as a comprehensive approach through the integration of plans and strategies


related to climate change and sustainable agriculture has taken an effective step in
the direction of preserving natural resources and improving sustainable human ben-
efits. The international climate organizations have fully included the CSA frame-
work in their agenda, and policy-makers are often looking for ways to reduce
greenhouse gas emissions and adapt agricultural systems to climate change.
However, agricultural and food systems are complex systems with many agronomic,
sociopolitical, and economic interdependencies. In many cases, progress in one area
may negatively affect another and vice versa. Therefore, a major challenge for the
implementation of CSA in different geographical areas is related to the coordination
of policies and programs of this approach. In fact, it is CSA policies that recognize
these trade-offs and allow prioritization or reconciliation between the three goals
(reducing greenhouse gas emissions, increasing adaptation, and productivity) even
when there are conflicts among the strategies (Lewis & Rudnick, 2019).
CSA policies, programs, and strategies at national and regional levels can be dif-
ferent in each country. In general, the theory of climate change Assessment of the
Climate Smart Agriculture Strategy (EASAC) states that CSA strategies can be
implemented from four paths, including the policy, institutional, financial, and com-
munication paths. The results or what is expected to happen through the use of
EASAC by governments, academics, international collaborations, etc. can be evalu-
ated in each path at three levels: regional, national, and sub-national (Collazos
et al., 2021).
194 F. Fathi et al.

9.1 Policy Path

One of the necessary measures for the development of CSA is policy-making and
the creation of a legal framework for greater coordination of agricultural, environ-
mental, and food security policies with climate change. Active policy-making
requires coordination in all policy domains. This interaction is generally achieved
through the interaction between all relevant ministries to facilitate, formulate, and
implement policies and investment plans to deal with climate change. There is an
assessment to achieve CSA. Policy formulation has a direct result and policy imple-
mentation/evaluation has indirect results for CSA. The underlying hypothesis in the
policy direction is that CSA policies are generally formulated at the regional level.
In other words, only some reinforcements are made in the formulation of and imple-
mentation policies at the national level. In the formulation stage, new policies are
developed or adjustments are made in the existing policies. Then these policies
should be implemented (through programs and projects). In general, in some cases,
adjusting the existing policies and ensuring their implementation for the develop-
ment of CSA can be a great approach that can be used instead of the approach of
developing new policies. New policies can potentially be used when there are deep
policy gaps or existing policies are inadequate. If necessary, governments can use
financial or regulatory instruments in cases where there are policy gaps (FAO, 2019).
Another important aspect of policy-making in development of CSA is the use of
science. The relationship between science and policy is considered as an important
aspect of research efforts in the field of climate change and environment at the world
level. Science-policy engagement is a pragmatic way to increase the impact of sci-
entific research on policy-making, which involves joint interaction between
researchers and policy-makers through participation (demand-based collaborative
research processes), evidence (creating science credibility by adopting flexible
methods), and expansion (effective communication and capacity building) (Dinesh
et al., 2018). This interaction is used by ensuring that the scientific findings from
agricultural research can accelerate the actions of governments, the private sector,
and nongovernmental organizations (NGOs) for the development of CSA. In rela-
tion to the field of climate change, issues such as causes, effects, interplay of strate-
gies, time horizons, and changes in governance have complicated the science-policy
interaction (Lemos & Agrawal, 2006; Meadowcroft, 2007). It should be noted that
there are various policies in the field of agriculture that can be considered as tools
for the development of CSA. Some of these policies are as follows:
11  Policy Framework to Introduce Climate-Smart Agriculture 195

9.1.1 Economic Policies

Subsidies and Taxes on Inputs and Outputs

Supportive policies are among the most important economic policies in the agricul-
tural sector of developed and developing countries in the world. The methods of
supporting farmers are divided into three categories: price supports (such as guaran-
teed purchasing), input supports (subsidizing of chemical fertilizer, energy, seeds,
and credits), and other supports (payments for drought compensation and part of
agricultural product insurance premium). By providing incentives and removing
barriers to adoption, governments can create an enabling environment for producers
to adopt CSA practices. To create this environment and develop CSA, various stake-
holders such as farmers, private sector, financial institutions, and civil society have
an important role. It is essential for governments to align their policies and regula-
tions with the objectives of the CSA. If necessary, they should design new coherent
policies, strategies, plans, and programs for the development of CSA or adjust and
modify existing policies (Vermeulen et al., 2013).
Input support in the agricultural sector (e.g., subsidizing fertilizers, pesticides,
and seeds) may reduce farmer production costs and lead to increased income and
food security in low-income countries, but on the other hand, it can result in exces-
sive consumption of chemical fertilizers and pesticides that lead to water and soil
pollution. Determining whether a particular policy instrument supports efforts to
achieve CSA goals is therefore not always straightforward. Energy and water sub-
sidy is another type of support. Subsidies for energy or water consumption help
production, increase farmers’ income, and reduce their food security. But it also
leads to unsustainable actions. As an example, the payment of energy subsidies in
Iran has led to many economic problems. The payment of energy subsidies has led
to an increase in consumption by producers in various sectors, especially agricul-
ture, as a result of which the emission of greenhouse gases has also increased.
Eliminating these subsidies indirectly requires the development of CSA activities.
However, one of the effects of removing the energy subsidy is the reduction of
greenhouse gas emissions. The removal of energy subsidy has led to an increase in
the price of energy for consumers and an increase in energy efficiency, which can
affect the development process of CSA policies (Fathi & Bakhshoodeh, 2021).
Another example in this field is the payment of electricity and fuel subsidies for
agricultural water wells. This has led to increased extraction of groundwater
resources and inefficient use of water and has resulted in the degradation of ground-
water resources in India (Ignaciuk, 2015). Inadequate water pricing regulations dis-
incentivize water conservation. Additionally, this problem has led to increased
salinity in a number of irrigation-dependent economies (such as Pakistan and
Uzbekistan). In Egypt, farmers receive irrigation water for free, but indirectly pay
the government for this service. In order to finance irrigation systems, the Egyptian
government buys cotton from farmers at a fixed price, which is lower than the price
in the world market. On the other hand, the government earns money by selling cot-
ton in international markets from the difference in price. Finally, these revenues are
196 F. Fathi et al.

used to finance the irrigation system. A similar situation prevails in Uzbekistan


(Jumaboev et al., 2010). As another example, it can be mentioned that the use of
incentive policies in California has continuously improved the efficiency of water
delivery and irrigation systems over the past 50 years. These policies have led to the
development and improvement of infrastructure and the creation of drip and micro-­
rain irrigation systems in California, which has resulted in an increase in economic
productivity per unit of water consumed. Research has shown that using more effi-
cient irrigation methods and replacing flooded systems with subsurface drip sys-
tems can significantly reduce nitrogen oxide emissions from the soil. This is one of
the examples of CSA development.
Reducing the use of chemical fertilizers in agriculture and replacing them with
organic fertilizers is one of the necessities that can lead to the reduction of environ-
mental pollution. Providing subsidies for organic and biological fertilizers to man-
age the reduction of water and soil pollution can be effective in this field. Also,
governments may use the policy of taxing inputs and products to prevent activities
that pollute the environment. For example, in many developed countries, govern-
ments have increased the level of taxes on diesel used for agricultural production.
Recently, Austria and the Netherlands have also increased fuel taxes for farmers.
In general, environmental taxes are divided into two categories: direct and indi-
rect. Direct tax is the Pigovian tax that has a certain rate and is considered for each
unit of pollutant emission or environmental destruction. However, indirect taxes are
imposed on production inputs or consumer goods whose use leads to damage to the
environment. Carbon tax and energy rate tax are indirect environmental taxes.
Indirect environmental taxes such as carbon tax and energy price tax cause indirect
effects such as climate change (Nie et al., 2022). Nevertheless, one of the important
environmental threats related to energy is the issue of climate change, which is
closely related to the emission of greenhouse gases caused by energy production
and consumption. Various studies have been conducted on the effects of these types
of taxes on carbon dioxide emissions. The study of Parry and Mylonas (2017) is one
of these studies, which states that increasing the carbon tax has reduced greenhouse
gas emissions. Some studies have examined both carbon tax and energy tax in tan-
dem. For example, Fraser and Waschik (2013), Zou et al. (2018), and Khastar et al.
(2020) have investigated the effectiveness of these policies for South Africa,
Australia, China, and Finland, respectively. These types of policies indirectly affect
the development of CSA by influencing the emission of greenhouse gases and as a
result of climate change.

Social Supports

One of the challenges of moving towards CSA is ensuring that adequate supports
are provided to vulnerable families and communities. In some regions, climate
change is expected to make poor communities vulnerable to risks and ignite food
insecurity. In such a situation, there will probably be an urgent need for social sup-
port measures to reduce hunger and poverty. An important challenge in this field is
11  Policy Framework to Introduce Climate-Smart Agriculture 197

to identify suitable options for better coordination of social supports and CSA inter-
ventions. Social support includes things such as cash or noncash assistance to vul-
nerable households, protection of vulnerable people against weather risks,
improvement of people’s social status, and respect for the rights of marginalized
people. The aim of these plans is to reduce poverty and the economic and social
vulnerability of households (FAO, 2019).

Market and Trade-Based Tools

Extreme climatic events can intensify price fluctuations. Nonetheless, there will be
changes in the production patterns and the risk of disruption in the local and regional
supply chain will also increase. The existence of price risk in the market of inputs
and outputs leads to the risk of farmers’ income (Fathi et al., 2020), which is an
obstacle to the adoption of CSA technologies. In many countries, agricultural and
food markets for some specialized products are incomplete or nonexistent. Linking
smallholder farmers to local, national, and regional markets improves farmers’
access to the inputs and knowledge needed to move towards CSA. In other words,
this work opens more channels for selling products, which can also reduce the price
risk by creating competition. Both input and product markets are important and the
development of these markets or their integration can be a solution for the develop-
ment of CSA. For example, creating a market for crops used in crop rotation can be
a major driver for the adoption of CSA systems that are resilient to climate change
impacts and reduce greenhouse gas emissions. For example, the supply of seeds
suitable for the climate and soil conditions of a region provides good economic
opportunities for producers to reach new markets. Extreme weather events make it
more difficult to make decisions about the type of technology required for produc-
tion, the area under cultivation of crops, or the time of planting and harvesting. In
addition, these events emphasize the sensitivity and necessity of integrated planning
for production and distribution, especially in the agricultural sector (Ahmadi &
Abdollahzadeh, 2019).
Integrated planning of production and distribution is a focal point in the supply
chain management of goods and services. Integrated planning of production and
distribution can be considered as a strategy for making coordinated and consistent
decisions in order to increase the competitiveness potential in the process of produc-
tion and distribution of goods and services to meet the needs. In addition, this
method of planning leads to customer satisfaction in terms of optimizing other sup-
ply chain objectives such as costs and profit. With access to technology and meteo-
rological data, this type of planning can be used for farmers. This can reduce the
fluctuations and income risk of farmers and facilitate the adoption of CSA technolo-
gies. In addition to the direct effects of climate change on primary production, these
changes can affect comparative advantages and trade flows. In other words, climate
change can potentially change food trade and the international competitiveness of
some producers, especially producers in developing countries (FAO, 2016b). To
reduce price volatility in the international market, governments have experimented
198 F. Fathi et al.

with various trade-based instruments such as tariffs, import and export restrictions,
price controls, intervention purchasing, quotas, and subsidies. All these tools have
an economic cost and many of them can also create unwanted consequences (FAO,
2016a). However, they can be a solution to reduce the price risk for farmers.

9.2 Institutional Path

Institutional path: Institutions create spaces for dialogue and participation in the
field of promoting CSA policies. Therefore, in this way, they can help strengthen the
weaknesses and strengths of CSA implementation. In addition, institutions can play
an important role in streamlining CSA programs by developing research and exten-
sion, formulating policies, and implementing and evaluating CSA programs. The
second regional priority of CSA is to strengthen national and local institutions. This
priority is to support and increase the capacity to adapt to climate change through
the development of people’s access to statistics and information. The development
of these institutions has been considered as an effective strategy for a long time, but
they also require appropriate measures and financing. Empirical evidence shows
that institutions play a positive role in coordinating efforts that increase public inter-
est through the expansion and dissemination of information (Vermeulen et al., 2013).

9.3 Financial Path

Linking financial resources dedicated to combating climate change with conven-


tional agricultural resources is an important part of efforts to achieve CSA. The
purpose of creating a financial pathway is to support and facilitate access to finan-
cial resources to strengthen and promote CSA. Inadequate access to financial instru-
ments such as credit services is one of the biggest obstacles to the development of
CSA. Each country has its own plan and annual budget, and the allocation of funds
for CSA financing is different. In order to achieve CSA goals, it is necessary to align
appropriate funding sources with CSA implementation organizations. However,
financing, in addition to the investment budget and executive and operational bud-
get, also includes the necessary budget for maintenance and continuity of the proj-
ects. Although it is mostly annual and mid-term planning, adapting agricultural
systems to CSA requires an increase in initial investment, which is created by
increasing financial resources. This path is updated through the development of a
series of CSA projects and receiving financial assistance from international and
national funds, the Global Environment Facility (GEF), and the Green Climate
Fund (GCF). In many developing countries, long-term financing for smallholder
agriculture is not feasible. However, short-term loans and microcredit facilities may
be a good option. For example, studies in southern Africa have shown that only 16%
11  Policy Framework to Introduce Climate-Smart Agriculture 199

of agricultural households in Malawi and 10% in Zambia receive credit from formal
and informal sources.
Although improving access to long-term financing is important for replacing old
trees with newer and higher-yielding species, it may result in reduced income for
farmers in the short term. Therefore, access to long-term financial resources can
prevent the risk of reducing farmers’ income. In general, medium- and long-term
financing is important to support initial investment in technologies. Nonetheless,
the use of new technologies and innovations, in the long run, ends up increasing
productivity, improving the efficiency in the use of resources, flexibility against
climate shocks, and reducing the emission of pollution. In addition to directing
investments and financial aid towards CSA, it is recommended that other budget
decisions and investments in agricultural sectors be climate-proof. With this
approach, investment plans are evaluated through a climate change lens and poten-
tial climate changes or greenhouse gas emissions are monitored. For example, the
FAO Ex Ante Carbon-balance Tool (Ex Act) is a widely used assessment system
that provides estimates of the impact of agricultural and forestry development proj-
ects, programs, and policies on the carbon balance. Despite the importance of finan-
cial support, it should be emphasized that financial support through specific policy
measures is not enough to implement CSA (Inman et al., 2018). However, nonfinan-
cial aspects such as technical, managerial, and political considerations may create
obstacles in the way of realizing this goal (Giordano et al., 2015). Therefore, in the
context of adopting programs that can help the development of CSA, in addition to
financing, economic, and social factors, agricultural structure, and attitudes and
motivations of farmers should also be taken into account (Li et al., 2019; Price &
Leviston, 2014).

9.4 Communication Path

This path focuses on developing communications at the regional and national levels
to raise awareness and strengthen the implementation of EASAC. Given the limited
amount of government resources available, it is necessary for governments to
engage with the private sector. This helps to provide parts of long-term and cheap
credits for producers. Also, through this, farmers can also invest in the development
of CSA systems.

10 Conclusions

The main aim of this chapter was to develop a policy framework to introduce
CSA. The four policies, institutional, financing, and communication pathways out-
lined in this chapter, can help local farmers adopt CSA practices, services, and tech-
nologies (Fig. 11.1). This can facilitate the achievement of more efficient, adaptable,
200 F. Fathi et al.

Fig. 11.1  The framework of introducing CSA

and flexible agriculture in the face of climate change. In other words, these four
pathways are the fundamentals of introducing CSA. In fact, efficient and effective
operation of the CSA elements including CSA techniques, practices, methods, inno-
vations, interactions, arrangements, etc. depends on the existence of these four fun-
damental requirements. The use of this framework can provide a basis for easier
introduction of CSA in agricultural societies, reduction of greenhouse gas emis-
sions, and mitigation of the effects of climate change.

References

Abegunde, V. O., Sibanda, M., & Obi, A. (2019). Determinants of the adoption of climate-smart
agricultural practices by small-scale farming households in King Cetshwayo District munici-
pality, South Africa. Sustainability, 12(1), 195. https://doi.org/10.3390/su12010195
Abera, W., Assen, M., & Budds, J. (2020). Determinants of agricultural land management practices
among smallholder farmers in the Wanka watershed, northwestern highlands of Ethiopia. Land
Use Policy, 99, 104841. https://doi.org/10.1016/j.landusepol.2020.104841
Adesipo, A., Fadeyi, O., Kuca, K., Krejcar, O., Maresova, P., Selamat, A., & Adenola, M. (2020).
Smart and climate-smart agricultural trends as core aspects of smart village functions. Sensors,
20(21), 5977. https://doi.org/10.3390/s20215977
Ahmadi, K., & Abdollahzadeh, S. (2019). A model for the integration of production-distribution
levels in the supply chain of nonperishable materials by considering intermediate warehouses.
Journal of Production and Operations Management, 10(2), 37–53.
11  Policy Framework to Introduce Climate-Smart Agriculture 201

Ali, A., & Erenstein, O. (2017). Assessing farmer use of climate change adaptation practices and
impacts on food security and poverty in Pakistan. Climate Risk Management, 16, 183–194.
https://doi.org/10.1016/j.crm.2016.12.001
Ashraf, I., Ahmad, I., Nafees, M., Yousaf, M. M., & Ahmad, B. (2021). A review on organic farm-
ing for sustainable agricultural production. Pure and Applied Biology, 5(2), 277–286. https://
doi.org/10.19045/bspab.2016.50036
Aweke, M. (2017). Gelaw climate-smart agriculture in Ethiopia: CSA country profiles for Africa
series. International Center for Tropical Agriculture.
Azadi, H., Movahhed Moghaddam, S.  M., Burkart, S., Mahmoudi, H., Van Passel, S., Kurban,
A., & Lopez-Carr, D. (2021). Rethinking resilient agriculture: From climate-smart agricul-
ture to vulnerable-smart agriculture. Journal of Cleaner Production, 319, 128602. https://doi.
org/10.1016/j.jclepro.2021.128602
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting cir-cular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Bhattacharyya, P., Pathak, H., & Pal, S. (2020). Climate smart agriculture: Concepts, challenges,
and opportunities. Springer.
Branca, G., Tennigkeit, T., Mann, W., & Lipper, L. (2012). Identifying opportunities for climate-­
smart agriculture investment in Africa (p.  132). Food and Agriculture Organization of the
United Nations.
Bryan, E., Ringler, C., Okoba, B., Koo, J., Herrero, M., & Silvestri, S. (2013). Can agriculture
support climate change adaptation, greenhouse gas mitigation and rural livelihoods?Insights
from Kenya. Climatic Change, 118(2), 151–165. https://doi.org/10.1007/s10584-­012-­0640-­0
Campbell, B.  M., Thornton, P., Zougmoré, R., Van Asten, P., & Lipper, L. (2014). Sustainable
intensification: What is its role in climate smart agriculture? Current Opinion in Environmental
Sustainability, 8, 39–43. https://doi.org/10.1016/j.cosust.2014.07.002
Chandra, A., McNamara, K. E., & Dargusch, P. (2018). Climate-smart agriculture: Perspectives
and framings. Climate Policy, 18(4), 526–541. https://doi.org/10.1080/14693062.2017.131696
8
Collazos, S., Howland, F., & Le Coq, J.  F. (2021). Towards a theory-based assessment of the
Climate Smart Agriculture Strategy (EASAC) for the SICA Region.
Deressa, T. T., Ringler, C., & Hassan, R. M. (2010). Factors affecting the choices of coping strate-
gies for climate extremes. The case of farmers in the Nile Basin of Ethiopia IFPRI Discussion
Paper, 1032.
Deresse, M., & Zerihun, A. (2018). Financing challenges of smallholder farmers: A study on mem-
bers of agricultural cooperatives in Southwest Oromia region, Ethiopia. African Journal of
Business Management, 12(10), 285–293. https://doi.org/10.5897/AJBM2018.8517
Dinesh, D., Zougmore, R. B., Vervoort, J., Totin, E., Thornton, P. K., Solomon, D., Shirsath, P.,
Pede, V., Lopez Noriega, I., Läderach, P., Körner, J., Hegger, D., Girvetz, E., Friis, A., Driessen,
202 F. Fathi et al.

P., & Campbell, B.  M. (2018). Facilitating change for climate-smart agriculture through
science-­policy engagement. Sustainability, 10(8), 2616. https://doi.org/10.3390/su10082616
Engel, S., & Muller, A. (2016). Payments for environmental services to promote “climate-smart
agriculture”? Potential and challenges. Agricultural Economics, 47(S1), 173–184. https://doi.
org/10.1111/agec.12307
Ewbank, R. (2015). Climate-resilient agriculture: What small-scale producers need to adapt to
climate change. Time Climate Justice, 15, 2016–2003.
Fathi, F., & Bakhshoodeh, M. (2021). Economic and environmental strategies against targeting
energy subsidy in Iranian meat market: A game theory approach. Energy Policy, 150, 112153.
https://doi.org/10.1016/j.enpol.2021.112153
Fathi, F., Sheikhzeinoddin, A., & Talebnejad, R. (2020). Environmental and economic risk man-
agement of seed maize production in Iran. Journal of Cleaner Production, 258, 120772. https://
doi.org/10.1016/j.jclepro.2020.120772
Food and Agriculture Organization. (2010). Climate-smart agriculture: Policies, practices and
financing for food security, adaptation and mitigation. Author.
Food and Agriculture Organization. (2014). IWMI. Wetland and agriculture: Partners; the Ramsar
Convention on wetlands.
Food and Agriculture Organization. (2016a). The state of food and agriculture. Climate change,
agriculture and food security.
Food and Agriculture Organization. (2016b). Climate change and food security: Risks and
responses.
Food and Agriculture Organization. (2019). Enabling policy environment for climate-smart agri-
culture. Food and Agriculture Organization.
Food and Agriculture Organization (FAO) of the United Nations. (2013). Climate-smart
agriculture-­sourcebook. E-ISBN978-92-5-107721-4.
Fraser, I., & Waschik, R. (2013). The double dividend hypothesis in a CGE model: Specific
factors and the carbon base. Energy Economics, 39, 283–295. https://doi.org/10.1016/j.
eneco.2013.05.009
García de Jalón, S.  G., Silvestri, S., & Barnes, A.  P. (2017). The potential for adoption of cli-
mate smart agricultural practices in sub-Saharan livestock systems. Regional Environmental
Change, 17(2), 399–410. https://doi.org/10.1007/s10113-­016-­1026-­z
Giordano, R., D’Agostino, D., Apollonio, C., Scardigno, A., Pagano, A., Portoghese, I.,
Lamaddalena, N., Piccinni, A. F., & Vurro, M. (2015). Evaluating acceptability of groundwa-
ter protection measures under different agricultural policies. Agricultural Water Management,
147, 54–66. https://doi.org/10.1016/j.agwat.2014.07.023
Gupta, J. (2009). Climate change and development cooperation: Trends and questions.
Current Opinion in Environmental Sustainability, 1(2), 207–213. https://doi.org/10.1016/j.
cosust.2009.10.004
Gupta, J. (2010). A history of international climate change policy. WIREs Climate Change, 1(5),
636–653. https://doi.org/10.1002/wcc.67
Harvey, C. A., Chacón, M., Donatti, C. I., Garen, E., Hannah, L., Andrade, A., Bede, L., Brown,
D., Calle, A., Chará, J., Clement, C., Gray, E., Hoang, M. H., Minang, P., Rodríguez, A. M.,
Seeberg-Elverfeldt, C., Semroc, B., Shames, S., Smukler, S., et al. (2014). Climate-smart land-
scapes: Opportunities and challenges for integrating adaptation and mitigation in tropical agri-
culture. Conservation Letters, 7(2), 77–90. https://doi.org/10.1111/conl.12066
Hellin, J., & Fisher, E. (2019). Climate-smart agriculture and non-agricultural livelihood transfor-
mation. Climate, 7(4), 48. https://doi.org/10.3390/cli7040048
Ignaciuk, A. (2015). Adapting agriculture to climate change: A role for public policies.
Inman, A., Winter, M., Wheeler, R., Vrain, E., Lovett, A., Collins, A., Jones, I., Johnes, P., &
Cleasby, W. (2018). An exploration of individual, social and material factors influencing water
pollution mitigation behaviors within the farming community. Land Use Policy, 70, 16–26.
https://doi.org/10.1016/j.landusepol.2017.09.042
11  Policy Framework to Introduce Climate-Smart Agriculture 203

Josephson, A. L., Ricker-Gilbert, J., & Florax, R. J. G. M. (2014). How does population density
influence agricultural intensification and productivity? Evidence from Ethiopia. Food Policy,
48, 142–152. https://doi.org/10.1016/j.foodpol.2014.03.004
Jumaboev, K., Eshmuratov, D., Reddy, J. M., Anarbekov, O., & Kazbekov, J. S. (2010). Prediction
of improved water productivity on-farm level in the selected cotton farms of Fergana and
Andijan Provinces of Uzbekistan.
Khastar, M., Aslani, A., & Nejati, M. (2020). How does carbon tax affect social welfare and emission
reduction in Finland? Energy Reports, 6, 736–744. https://doi.org/10.1016/j.egyr.2020.03.001
Lemos, M. C., & Agrawal, A. (2006). Environmental governance. Annual Review of Environment
and Resources, 31(1), 297–325. https://doi.org/10.1146/annurev.energy.31.042605.135621
Lewis, J., & Rudnick, J. (2019). The policy enabling environment for climate smart agriculture: A
case study of California. Frontiers in Sustainable Food Systems, 3, 31. https://doi.org/10.3389/
fsufs.2019.00031
Li, P., Chen, Y., Hu, W., Li, X., Yu, Z., & Liu, Y. (2019). Possibilities and requirements for intro-
ducing agri-environment measures in land consolidation projects in China, evidence from eco-
system services and farmers’ attitudes. Science of the Total Environment, 650(2), 3145–3155.
https://doi.org/10.1016/j.scitotenv.2018.10.051
Lima, M.  G. B. (2016). Policies and practices for climate-smart agriculture in Sub-Saharan
Africa: A comparative assessment of challenges and opportunities across 15 countries. Food,
Agriculture, and Natural Resources Policy Analysis Network (FANRPAN).
Lipper, L., & Zilberman, D. (2018). A short history of the evolution of the climate smart agricul-
ture approach and its links to climate change and sustainable agriculture debates (pp. 13–30).
Springer. https://doi.org/10.1007/978-­3-­319-­61194-­5_2.
Lipper, L., Thornton, P., Campbell, B.  M., Baedeker, T., Braimoh, A., Bwalya, M., Caron, P.,
Cattaneo, A., Garrity, D., Henry, K., Hottle, R., Jackson, L., Jarvis, A., Kossam, F., Mann, W.,
McCarthy, N., Meybeck, A., Neufeldt, H., Remington, T., et al. (2014). Climate-smart agricul-
ture for food security. National Climatic Change, 4, 1068–1072.
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In Climate change:
The social and scientific construct by Suhaib A. Bandh. Springer. Nature Publishing. https://
link.springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Mango, N., Makate, C., Mapemba, L., & Sopo, M. (2018). The role of crop diversification in
improving household food security in Central Malawi. Agriculture and Food Security,
7(1), 1–10.
Meadowcroft, J. (2007). Who is in charge here? Governance for sustainable development in a
complex world. Journal of Environmental Policy and Planning, 9(3–4), 299–314. https://doi.
org/10.1080/15239080701631544
Mohammed, E. (2016). Opportunities and challenges for adopting conservation agriculture at
smallholder Farmer’s level the case of Emba Alage, Tigray, Northern Ethiopia (Doctoral
Dissertation). St. Mary’s University.
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Mwongera, C., Shikuku, K. M., Twyman, J., Laderach, P., Ampaire, E., Van Asten, P., Twomlow,
S., & Winowiecki, L. (2017). Climate smart agriculture rapid appraisal (CSARA): a tool for
prioritizing context-specific climate smart agriculture technologies. Agricultural Systems, 151,
192–203.
Nagothu, U.  S., & Kolberg, S. (2016). Climate smart agriculture: Is this the new paradigm of
agricultural development? In Climate change and agricultural development (pp.  17–36).
Routledge.
Nie, P. Y., Wang, C., & Wen, H. X. (2022). Optimal tax selection under monopoly: Emission tax vs
carbon tax. Environmental Science and Pollution Research International, 29(8), 12157–12163.
https://doi.org/10.1007/s11356-­021-­16519-­1
204 F. Fathi et al.

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. https://doi.org/10.1016/j.wasman.2015.07.038
Ogato, G. S. (2014). Biophysical, socio-economic, and institutional constraints for production and
flow of cereals in Ethiopia. American. Journal of Human Ecology, 3(3), 51–71. https://doi.
org/10.11634/216796221403571
Ojoko, E. A., Akinwunmi, J. A., Yusuf, S. A., & Oni, O. A. (2017). Factors influencing the level
of use of climate-smart agricultural practices (CSAPs) in Sokoto state, Nigeria. Journal of
Agricultural Sciences, Belgrade, 62(3), 315–327. https://doi.org/10.2298/JAS1703315O
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Parry, I. W. H., & Mylonas, V. (2017). Canada’s carbon price floor. National Tax Journal, 70(4),
879–900. https://doi.org/10.17310/ntj.2017.4.09
Partey, S.  T., Zougmoré, R.  B., Partey, S.  T., Ouédraogo, M., Torquebiau, E., & Campbell,
B. M. (2018). Facing climate variability in sub-Saharan Africa: Analysis of climate-smart agri-
culture opportunities to manage climate-related risks. Cahiers Agricultures (TSI), 27(3), 1–9.
Price, J. C., & Leviston, Z. (2014). Predicting pro-environmental agricultural practices: The social,
psychological and contextual influences on land management. Journal of Rural Studies, 34,
65–78. https://doi.org/10.1016/j.jrurstud.2013.10.001
Rosenstock, T.  S., Lamanna, C., Chesterman, S., Bell, P., Arslan, A., Richards, M., Riou, J.,
Akinleye, A. O., Champalle, C., Cheng, Z., Corner-Dolloff, C., Dohn, J., English, W., Eyrich,
A.  S., Girvetz, E.  H., Kerr, A., Lizarazo, M., Madalinska, A., McFatridge, S., . . . Strom,
H. (2016). The scientific basis of climate-smart agriculture: A systematic review protocol.
CCAFS Working Paper no. 138.
Seebauer, M. (2014). Whole farm quantification of GHG emissions within smallholder
farms in developing countries. Environmental Research Letters, 9(3), 035006. https://doi.
org/10.1088/1748-­9326/9/3/035006
Subedi, R., Bhatta, L. D., Udas, E., Agrawal, N. K., Joshi, K. D., & Panday, D. (2019). Climate-­
smart practices for improvement of crop yields in mid-hills of Nepal. Cogent Food and
Agriculture, 5(1), 1631026. https://doi.org/10.1080/23311932.2019.1631026
Tan, X. (2010). Clean technology R&D and innovation in emerging countries—Experience from
China. Energy Policy, 38(6), 2916–2926. https://doi.org/10.1016/j.enpol.2010.01.025
Thorn, J.  P., Friedman, R., Benz, D., Willis, K.  J., & Petrokofsky, G. (2016). What evidence
exists for the effectiveness of on-farm conservation land management strategies for preserv-
ing ecosystem services in developing countries? A systematic map. Environmental Evidence,
5(1), 1–29.
Vaast, P., Harmand, J. M., Rapidel, B., & Jagoret, P. (2016). Deheuvels, O. coffee and cocoa pro-
duction in agroforestry—A climate-smart agriculture model. In Climate change and agricul-
ture worldwide (pp. 209–224). Springer.
Vermeulen, S.  J., Challinor, A.  J., Thornton, P.  K., Campbell, B.  M., Eriyagama, N., Vervoort,
J.  M., Kinyangi, J., Jarvis, A., Läderach, P., Ramirez-Villegas, J., Nicklin, K.  J., Hawkins,
E., & Smith, D.  R. (2013). Addressing uncertainty in adaptation planning for agriculture.
Proceedings of the National Academy of Sciences of the United States of America, 110(21),
8357–8362. https://doi.org/10.1073/pnas.1219441110
Woolf, D., Solomon, D., & Lehmann, J. (2018). Land restoration in food security programmes:
Synergies with climate change mitigation. Climate Policy, 18(10), 1260–1270. https://doi.org/1
0.1080/14693062.2018.1427537
Zeng, D., Alwang, J., Norton, G., Jaleta, M., Shiferaw, B., & Yirga, C. (2018). Land ownership
and technology adoption revisited: Improved maize varieties in Ethiopia. Land Use Policy, 72,
270–279. https://doi.org/10.1016/j.landusepol.2017.12.047
11  Policy Framework to Introduce Climate-Smart Agriculture 205

Zerga, B., & Gebeyehu, G. (2016). Climate change in Ethiopia variability, impact, mitigation, and
adaptation. Journal of Social Science and Humanities Research, 2(4), 66–84.
Zerssa, G., Feyssa, D., Kim, D. G., & Eichler-Löbermann, B. (2021). Challenges of smallholder
farming in Ethiopia and opportunities by adopting climate-smart agriculture. Agriculture,
11(3), 192.
Zou, L., Xue, J., Fox, A., & Meng, B. (2018). The emissions reduction effect and economic impact
of an energy tax vs. a carbon tax in China: A dynamic cge model analysis. Singapore Economic
Review, 63(2), 339–387. https://doi.org/10.1142/S021759081740015X
Chapter 12
Technological and Managerial Innovation
in Agriculture to Ensure Food Security
Under Climate Change

Fouad Elame, Hayat Lionboui, and Mohammed Behnassi

Abstract  Agriculture is still the most critical sector for food security, especially in
resource-constrained countries from the Global South. Indeed, this sector is increas-
ingly invested with the key purpose to provide healthy, safe, and nutritious food for
a growing food demand. However, the current challenges consist of meeting this
objective while sustainably using natural resources, conserving biodiversity, reduc-
ing emissions and pollution, and adapting to climate change. To meet these chal-
lenges and seize related opportunities, the agriculture sector needs to undergo a
transformative change through innovation. The integration of new technologies and
good management practices in this sector are response mechanisms that have the
potential to help attain both sustainability and food security. By reference to the
Moroccan context, marked by the strategic importance of agriculture for both devel-
opment and security of the country, this chapter aims to analyze the agricultural
sector within the perspective of identifying the areas where the technological and
managerial innovation may boost productivity while ensuring sustainability and
adaptation to climate change.

Keywords  Agriculture · Food security · Climate change · Natural resources ·


Technological innovation

F. Elame (*)
INRA, Regional Center of Agronomic Research of Agadir, Agadir, Morocco
e-mail: fouad.elame@inra.ma
H. Lionboui
INRA, Regional Center of Agronomic Research of Rabat, Rabat, Morocco
M. Behnassi
University Ibn Zohr, Agadir, Morocco

© The Author(s), under exclusive license to Springer Nature 207


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_12
208 F. Elame et al.

1 Introduction

Facing a growing population and changing conditions of production systems, con-


sumption, and trade, food supply has become a major issue, renewed by the food
crisis of 2008 (Heady & Fan, 2011). This food crisis has shown the limits of the
current agricultural development and innovation model promoted at the interna-
tional level (Allaire & Daviron, 2017). Food security is mostly linked to the avail-
ability, quality (especially health and nutritious), regularity, and accessibility of
food. Globally, around 75% of families, below the poverty line, are living in rural
areas and depending on farming activities (World Bank, 2016). According to the
FAO (2015), around 795 million people, or 1 in 9 people, are undernourished.
Developments in agriculture and food cause rapid damage of natural resources.
The question first concerns biodiversity, which is heavily impacted by deforestation,
agricultural intensification, and the overuse of chemical products by agriculture.
The issue also concerns the management of soils, damaged by agricultural activities
(erosion, loss of fertility, etc.), and that of water, widely used for irrigation and pol-
luted by pesticide residues and fertilizers. Agricultural and related agri-food activi-
ties are also at the core of the nitrogen and phosphorus cycle crisis (Rockström
et al., 2009). Nowadays, their negative impacts even threaten agriculture production
systems in many areas (OCDE-FAO, 2016). Agriculture and food then appear at the
center of the environmental debate because they are one of the causes of the degra-
dation observed, but also because they are first and foremost part of the solutions
(Fuglie et al., 2020).
Integrated natural resource management, poverty reduction, and capacity build-
ing are also major challenges facing agricultural and food activities, primarily in the
least developed countries, where the rural population is large and the poorest one.

2 Agriculture in the Process of Transformation: Current


State and Development Prospects

Since the independence, Morocco has admittedly made efforts to improve the food
balance by giving priority to agriculture and hydro-agricultural development.
However, despite these efforts, the food balance remains in deficit for several basic
products. According to the FAO, the cereal import dependency ratio was 54% in
2011/12 and 42% in 2016; the sugar import dependency ratio increased from an
annual average of 49% between 1992 and 2008 to 71% between 2009 and 2016. The
high dependence ratio on oil imports also increased, from 66% on average in
1986–1987 to 90% in 2000–2001, reaching 98.5% in 2009–2010 (FAO-OCDE,
2017). In addition, the competitiveness and sustainability of Moroccan agriculture
suffer from some constraints and they are eventually threatened by the scarcity of
land and water resources and the impacts of climate change (Malla et  al., 2022;
Bandh et al., 2021, 2022, 2023; Mushtaq et al., 2020; Parray et al., 2022; Bandh,
12  Technological and Managerial Innovation in Agriculture to Ensure Food Security… 209

2022a, b). In fact, there are many constraints facing the agricultural sector. Some are
related to the fragile character of resources and climate change impacts; others are
mainly the result of organizational and decision-making gaps and failures (Elame
et al., 2021).
If we only consider the impact of climate and land constraints in rainfed agricul-
ture areas, covering 90% of the useful agricultural land with a population of around
12 million inhabitants, who do not have such importance such as the irrigated sec-
tor, we can observe multiple adverse effects on the dynamics of agricultural devel-
opment and food security.
The rainfed areas are more marked by the existence of socioeconomic constraints
which manifest through the low level of equipment in these rainfed areas, analpha-
betism, poverty, and insufficient investments of the government.
The share of agriculture in the GDP formation was estimated at 13.6% in 2017
(HCP, cited by Harbouze et al., 2019). Improving the use of natural resources in
agriculture can effectively contribute to increasing production and food security.
Nowadays, agriculture must increase its production levels while limiting the use of
inputs and preserving the health and safety of consumers. To meet this challenge, a
profound change in practices is required, with a view of saving labor due to the lack
of attractiveness of the agricultural sector and the necessary maintenance of
competitiveness.
Agricultural development in Morocco is undergoing a transformation period.
The technological advances offer new perspectives. Indeed, the Moroccan agricul-
tural market represents an appealing market for technology used to generate infor-
mation at high frequency and analyzed to understand the changes and react
immediately. This underway digitalization slot comes in response to requests from
farmers who need real-time data to have a global view of their farm, which has
become a staple of the sustainable development strategy of many countries (Lionboui
et al., 2022).
This aspect of modern agriculture represents an important source of data and one
of the pillars of the “green generation” policy 2020–2030 launched by the Moroccan
Kingdom in 2020 which sets out the modernization strategy of the Moroccan agri-
cultural sector, by opening huge projects to boost this sector of the Moroccan econ-
omy which remains an attractive sector for young people and investors. For several
experts, smart agriculture is no longer a trend, but a real proof of development.
Accordingly, farmers are trying to adapt to new changes and changing their prac-
tices. In this sense, agriculture has become a field of permanent innovation in differ-
ent subjects, whether robotics, biocontrol and biotechnology, camera analysis, or
soil conservation using computer algorithms and which are accessible in real time
via applications and smartphones.
From this perspective, the current context shows that there is a real opportunity
to make agriculture interact with innovation and entrepreneurship. Indeed, the
Moroccan agricultural sector constitutes fertile ground for innovative project oppor-
tunities which will certainly change the current agricultural model. It is at this level
that entrepreneurship and innovative ideas appear as the common thread that can
lead to innovative agriculture. The retention of young people in rural areas is also an
210 F. Elame et al.

issue as most young people move towards jobs in urban areas. Technological inno-
vations can be an asset in attracting young people to rural areas by offering innova-
tive solutions for the Moroccan farm (Yatribi, 2020).

3 Innovation Technologies in Agriculture and Food System

In arid and semiarid areas, agriculture and food system performance depend more
on the weather and the possible impacts of climate change on productions should be
considered with most concern in areas where rainfed agriculture is still the main
source of income. Today, the challenges of climate change and food security are
pushing countries to integrate the new technology aspect to seek other levers for
agricultural development. Indeed, more and more efforts are deployed in ways to
connect farmers to improve the production value chain. According to the FAO report
of 2018 on technological innovation in agriculture, technology development and
technological innovations appear to be one of the important means of strengthening
competitiveness, growth, and employment and even of reducing social inequali-
ties (FAO, 2018).
Before developing this section, we need to start first defining some concepts.
Innovation could be defined as the development and introduction of a new idea and
transforming that idea into a product, process, object, or service (Şimşit et al., 2014).
O’Sullivan (2000) stated that “Innovation is the process through which productive
resources are developed and utilized to generate higher quality and/or lower cost
products than had previously been available.”
Innovation systems represent a collaborative system that aggregates multiple
organizations and individuals working to achieve specific goals or improvement.
Innovation could target different dimensions related to a product, a process, a mana-
gerial, and organizational method.
An innovation system integrates the invented system and the complementary
operations that will make it possible to transform the invention into innovation and
the ensuing dissemination and use.
Innovation technologies represent technologies and improvement created and
developed by innovation systems. In agriculture, it can be called agricultural inno-
vation system (AIS) (Fig. 12.1).
Nowadays, the introduction of technological innovations has now become an
imperative: on the one hand, to follow the global trend and align with competitors
and, on the other hand, to respond to structural problems in the sector (efficiency
and competitiveness issues, employment, social inequalities alleviation, etc.). These
are, in fact, new technological trends that are establishing themselves around the
world as essential difference tools for the competitiveness of actors in the agricul-
tural field. The literature clearly indicates the gap between innovative agriculture
and conventional agriculture based on the experience and judgment of the farmer.
Experts now agree that technological innovation is an asset to save time and inputs
12  Technological and Managerial Innovation in Agriculture to Ensure Food Security… 211

Institutional
analysis and
development
Agricultural
Innovation system
(Ideas and Knowledge
generation,
application and
diffusion)
Technology
development

Fig. 12.1  Innovation system framework in agriculture

for the interest of the grower, the community, and the environment (Faure et  al.,
2018; Feder et al., 1985; Rijswijk et al., 2019).
Modernizing the agricultural sector involves the introduction of technological
innovations at different stages of the production process, from upstream to down-
stream of the value chain. At the farm level, many innovations are related to pro-
cesses and practices since they consist in improving production techniques, for
example, by adopting improved seeds or more suitable irrigation systems.
Downstream industries are also innovating by making further improvements to their
products, as seen with the functional benefits embedded in food quality or the chem-
ical or pharmaceutical industry. Along the supply chain, organizational and market-
ing innovations are also very important.
Digital agriculture is identified as one of the ten key areas of digital technologies,
with three areas of focus: robotics, precision agriculture, and big data. Agriculture
is therefore seizing on these devices which produce a huge amount of data from
which knowledge and precise information could be extracted using data mining
techniques (Han et al., 2011). The use of new information and communication tech-
nologies in agriculture, first associated with precision farming, extends to a wide
number of innovations that combine sensors, robots, onboard computing, digital
herd or irrigation management applications and decision support tools, etc.
Consequently, the multiple applications forming the basis of connected agriculture
are changing the conditions of the agricultural work and the management of the
farm (Bellon-Maurel & Huyghe, 2016; Touzard, 2018).
The purpose of precision farming technologies is to optimize agricultural pro-
duction. They predict the outbreak of diseases in a plot and reduce fossil energy
consumption or the laboriousness of the working activities. They could also accu-
rately define and manage the water, fertilizer, and phytosanitary product require-
ments of crops. It thus becomes possible to optimize the use of chemical inputs and
equipments, as well as a reduction in CO2 emissions (FAO, 2008).
212 F. Elame et al.

Digital agriculture can also help to monitor plant health by triggering agricul-
tural warnings. In fact, pest and disease propagation in agriculture are linked to
climatic conditions, whereas by knowing these risks in advance, through numerical
forecasting, some diseases could be prevented and farmers could apply less chemi-
cal treatments and therefore better integrated protection management (IPM).
Innovation in machinery has been a very strong lever, with the appearance of
very original options, such as the management of weeds (weeds) by crushing their
seeds at the exit of the combine harvester. It is  an innovative method to achieve
integrated weed management and minimize herbicide use. The use of the automatic
tractor is one of the important technologies that saves a lot of time and provides a
better way of field management (Touzard, 2018).
Digitalization can also be used downstream of the value chain to reduce the num-
ber of speculators and deliver agricultural goods directly through “marketplaces,” as
well as blockchains which are very useful for food traceability throughout the pro-
duction and processing chain to the consumer. The development of e-commerce is
now increasingly affecting agricultural products and food industry. Online shopping
catalogs have tremendously increased for practical answers to a societal need for
food products without wasting time and we could mention here the “Amazon” expe-
rience as an example. New technologies allow the farmer and agro-industry to have
the right information and can scale their project without having to invest too much.
It is possible, through technology, to adapt the answers to the needs. In addition, a
new area of economic activity has been created through advances in storage, conser-
vation, and transportation technologies. This has allowed food companies to develop
marketing chains that deliver goods throughout the world at high speed and rela-
tively low cost.
As for the Moroccan context, we have to take into account climate conditions,
demographic dimension, and analphabetism; reducing waste and producing more
are crucial. A revolution is already underway in Morocco particularly due to digital
technology development. The main objective is to produce more with less input and
to guarantee food security. In this sense, technological and human development
must go in parallel. Digital technology is today an important opportunity for
Moroccan agriculture and which depends on many parameters: climate, natural
resources, technical management, seeds, and the use of pesticides. To better control
these parameters and to better take decisions at the right time, it is important to have
the right information, to be able to store it, and to analyze it accordingly, hence the
interest of digital for agriculture.
We can list three types of impacts related to technological development: the first
concerns productivity. One of the objectives is to improve profitability; for example,
solutions related to labor in agriculture are useful in order to have better visibility
on what is being done in the field and to improve the way of doing it through data
collection and storing. The second impact is cost optimization, through cost account-
ing on all field-level operations. For example, by reducing water consumption,
which is a big issue given that our country suffers from water stress due to drought
years, and also in terms of fertilizers use and all the costs related to equipment and
labor. The third impact is the improvement of quality with respect to the
12  Technological and Managerial Innovation in Agriculture to Ensure Food Security… 213

environment, in particular, the agricultural product quality related to the use of pes-
ticides which will allow respecting the export standards, knowing that the European
Union, the first commercial partner of Morocco, is putting more restrictions on
Moroccan agricultural products.
In this sense, technology can accelerate modernization, but at the same time can
create inequalities in terms of access to new technologies and sometimes noncom-
pliance with international standards. For this reason, regulations must accompany
this technological transformation. The innovation system is always subject to a cor-
pus of regulations supposed to guide it and to provide a global framework, taking
into account the interests of society and economic actors and ensuring fairness. This
regulation is translated into standards (for example, in animal genetics), national
catalogs, or intellectual property rules. In the case of plant genetics, in Europe, for
example, the varieties obtained are the subject of a study before inclusion in the
official national catalog. In France, innovations in the field of breeding technologies
are currently the subject of an assessment by the Scientific Committee of the
Permanent Technical Committee for Selection to determine how they are likely to
modify the variety supply and assessment processes.
Other restrictions related to the use of drones are facing the agricultural sector.
Drone technology seems to be very important for smart farming, particularly in
terms of data collection and processing, input, and time savings. Nevertheless, there
are many regulations and logistical challenges to overcome (Royer et al., 2020). The
most significant challenge to the widespread adoption of drones by the agricultural
industry is legal, as countries aim to preserve the legality of drone use while ensur-
ing aviation safety and protecting privacy rights.
Although some developed countries already have a detailed regulatory frame-
work, many other countries have not yet developed any regulations in this regard. In
Morocco, the acquisition of this technology is governed by an order bearing the
number 386-15 of the first article of the law number 13/89 of 2015. This law prohib-
its the import of drones without a license which must be specific to its use. In addi-
tion to this regulation, some farmers believe that their cost of acquisition is still high
compared to the results expected from such a technology.

4 The Agricultural Modernization in Morocco

This section highlights key technological and institutional innovations and key poli-
cies that might have an impact on the agricultural transformation. According to the
World Bank report (2020), investing in the adoption of new technologies could help
generate substantial gains in agricultural productivity and therefore in income. The
authors’ reports state  that nearly 80% of the extremely poor people are living in
rural areas, and most of them are practicing agriculture. Consequently, the fight
against poverty must be strongly focused on increasing agricultural productivity,
which has more impact than any other sector on poverty reduction (about twice as
much as the manufacturing industry). Developing countries need to increase
214 F. Elame et al.

Fig. 12.2 Climate-smart
agriculture approach
CLIMATE

SMART C.S.A AGRICULTURE


TECHNOLOGIES

. INCREASE PRODUCTIVITY AND INCOME

. BUILD RESILIENCE

. REDUCE GHG EMISSION

agricultural and food innovation and support the use of technology by farmers to
reduce poverty to meet the growing demand for food and overcome the negative
effects of climate change. Accordingly, a new concept was developed to achieve
these goals. Climate-smart-agriculture (CSA) is an innovative approach that inte-
grates the use of smart technologies in agriculture to reduce climate change impacts
(FAO, 2021). This approach incorporates adaptation, mitigation, and food security
and aims to manage the agricultural sector as the main component of food security
and increase its sustainably and productivity; build resilience and reduce the vulner-
ability of people and agri-food systems to climate change, and finally reduce GHG
emissions (Fig. 12.2).
Based on the CSA approach, Morocco has launched its new agricultural develop-
ment strategy, “Green Generation 2020–2030” which aims to improve the country’s
gross domestic product and double its exports over the next 10 years. This strategy
tends also to enhance human resources through the emergence of a new agricultural
middle class and a new generation of farmers by habilitating one million hectares of
collective land, creating 350,000 jobs and connecting at least two million farmers to
digital service platforms (ADA, 2020). Another objective is related to the improve-
ment of food distribution through the reorganization and modernization of whole-
sale markets.
In Morocco, modernization of the agricultural sector is marked by increased
adoption of modern technology and varieties and increased modern input use,
including chemical fertilizer and bio-fertilizers, machines, and more systematic
uses of water through drip irrigation technics.
12  Technological and Managerial Innovation in Agriculture to Ensure Food Security… 215

5 Research and Biotechnological Development

For varietal technology improvement, the contribution of the private sector in plant
breeding strategies is still very weak in Morocco. The public sector has developed
some varieties but is still very far from what is expected in comparison to the actual
biotechnological development in genetics and molecular applications at the interna-
tional level. Genetic improvement of some plant varieties can be used for nutrient
enrichment, improving tolerance to drought, herbicides, pests, and disease and
increasing yields per unit area or per plant given the problems of water scarcity and
land degradation (United Nations report, 2017). Strengthening intellectual property
rights is a key element in order to promote private-sector participation in varietal
development (Pray & Nagarajan, 2014).
Investing in research is also one of the weaknesses of the Moroccan agricultural
policy. Decision-makers must promote both public and private research and tech-
nology development and transfer to strengthen their overall innovation system. The
investments in the agricultural and food sector provide an opportunity to increase
economic growth, but the investment allocated to research remains very insufficient.
Investing in higher agricultural education and research and setting up strategies to
mobilize and involve the private sector is obligatory for any expected development.
The private sector can accelerate farmers’ access to new technologies. Private com-
panies contribute about half of total research-development (R&D) expenditure
focused on the needs of farmers in developed countries and up to a quarter in large
emerging economies such as China, India, and Brazil (UNTCAD, 2017). In terms
of policy and regulation, there are several ways to encourage increased private R&D
in agriculture: reducing barriers to market participation, setting up measures to pro-
mote competition, removing costly regulations, and strengthening intellectual prop-
erty rights.

6 Institutional and Policy Innovations

Morocco has seen various institutional reforms in the past few decades. The growth
of market institutions has been observed in the growing participation of new actors
in the processing, trading, and wholesale of agricultural and food products. Vertical
integration has also been strengthened for some crops like sugarcane and sugar beet.
However, constraints remain for food sanitation along the value chain for vegetables
and fruits and also for perishable products.
Production and marketing organization including aggregated farms and coopera-
tives have become increasingly important for the production and marketing of high-­
value vegetables and fruits, poultry, and higher-quality milk, which has led to the
organization of the market chain and the adoption of improved food safety prac-
tices. Indeed, the aggregation, which is one of the pillars of the Moroccan Green
Plan launched in 2008, is an innovative model for organizing farmers around private
216 F. Elame et al.

actors or professional organizations with good management knowledge. It is a win-


ning partnership between the productive upstream and the commercial downstream
of the market chain which helps overcoming the constraints linked to the fragmenta-
tion of land and the exorbitant cost of inputs.
This model is based on a contract established between aggregators holding agro-­
industrial development units and producers within the framework of cooperatives or
economic interest groups. Under this contract, the aggregator undertakes, in particu-
lar, to supervise the aggregated farmers and producers and to purchase their produc-
tion according to the terms defined in the aggregation contract. For their part, the
aggregated producers undertake to follow the technical itinerary recommended by
the aggregator and to deliver the agreed quantities. Consequently, this organiza-
tional model allows agro-industrials in particular to secure the supply of their units
with guaranteed and traceable quality production and allows aggregated producers
and farmers to benefit from modern production techniques and financing, to access
domestic and foreign market, to have greater bargaining power when negotiating
with buyers, and to obtain better access to inputs and services (FAO et al., 2017;
MAPMDREF, 2017).
According to many authors, the major weakness of contract farming is that it
may primarily benefit large farmers because transaction costs can be saved by mak-
ing a small number of contracts with large farmers rather than a large number of
contracts with smallholder farmers (Ton et al., 2018). In fact, the idea behind this
model was to perform the whole food and agricultural sector; unfortunately, the
main farms that benefit from all these great investments by the government are large
farms and small farmers who represent more than 70% (DEPF, 2019) did not benefit
much from this innovative model.
In addition to what has been developed in this section, we think that the process-
ing industry for agricultural products is not yet well developed in Morocco. Many
raw agricultural products are exported directly without any processing, which
means that the country loses a significant part of the added value. In this context,
agriculture and food industrialization of the country remains a key element to eco-
nomic development and to absorb the surplus labor force in the sector. Agribusiness
should be an important part of this industrialization, and without this key solution,
we will not assist to a real productivity improvement in the sector, no improvement
in the trade balance, and no reduction in unemployment even under this digitaliza-
tion strategy.

7 Conclusions

In a difficult economic context, technological development makes it possible to


offer new solutions for balanced socio-economic growth. Developing countries are
particularly vulnerable to problems related to climate change such as floods and
drought. Assistance in the form of research and capacity building that would enable
them to prevent and respond to growing problems is strongly needed. New
12  Technological and Managerial Innovation in Agriculture to Ensure Food Security… 217

technologies are opportunities in achieving the objectives of sustainable develop-


ment by decreasing input use and ensuring natural resource viability. Nowadays, we
are talking about “digital agriculture,” that is to say, agriculture that uses informa-
tion and communication technologies. These technologies can be implemented at
all agricultural production scales and other related fields, whether at the farm level
(optimization of crop production, herd management), in consulting services (new
agricultural advisory ways based on smart data use), or at larger scales such as in a
territory (water management optimization) or in a market chain (improvement of
inputs such as seeds, market information, etc.).
In the food and agriculture sector, adopting an integrated strategic approach rep-
resents the best way forward. The government has to invest in research and encour-
age the private sector for better management of available resources.
Finally, like all other economic activities, the agricultural sector is certainly
impacted by technological progress which could be useful for improving the socio-­
economic conditions of producers and populations.

References

ADA. (2020). Agence de Développement Agricole. Rapport de la Réunion régionale sur la nou-
velle stratégie “Génération Green 2020–2030”.
Allaire, G., & Daviron, B. (2017). Transformations agricoles et agroalimentaires: Entre écologie et
capitalisme, éditions Quæ, Versailles.
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting cir-cular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Bellon-Maurel, V., & Huyghe, C. (2016). L’innovation technologique dans l’agriculture.
Géoéconomie, 80(3), 159–180. https://doi.org/10.3917/geoec.080.0159
DEPF. (2019). Direction des études et des prévisions financières. Le secteur agricole marocain:
Tendances structurelles, enjeux et perspectives de développement. https://www.finances.gov.
ma/Publication/depf/2019/Le%20secteur%20agricole%20marocain.pdf
Elame, F., Lionboui, H., & Doukkali, R. (2021). The combined impact of climate change and
the use of solar energy on the water consumption in agriculture: A case study from souss
Massa region. Emerging challenges to food production and security in asia, middle east, and
Africa  – Climate risks and resource scarcity. Springer International Publishing. https://doi.
org/10.1007/978-­3-­030-­72987-­5.
218 F. Elame et al.

FAO. (2015). FAO, IFAD and WFP. The State of Food Insecurity in the World. Meeting the 2015
internationalhunger targets: taking stock of uneven progress. Rome. ISBN 978-92-5-108785-5.
FAO, C.-I., & CIRAD. (2017). étude sur l’agriculture familiale a petite échelle au proche-orient
et Afrique du nord pays focus maroc, 2017.
Faure, G., Chiffoleau, Y., Goulet, F., Temple, L., & Touzard, J. M. (2018). Innovation et développe-
ment dans les systèmes agricoles et alimentaires. Éditions Quæ. ISBN: 978-2-7592-2813-3.
Feder, G., Just, R. E., & Zilberman, D. (1985). Adoption of agricultural innovations in developing
countries: A survey. Economic Development and Cultural Change, 33(2), 255–298. Retrieved
February 10, 2021. http://www.jstor.org/stable/1153228. https://doi.org/10.1086/451461
Food and Agriculture Organization. (2008). Climate change and food security: A framework docu-
ment. FAO.
Food and Agriculture Organization. (2018). Des innovations technologiques ramènent les
jeunes vers l’agriculture: Comment les innovations technologiques stimulent l’emploi rural.
FAO, 2018.
Food and Agriculture Organization. (2021). Climate-smart agriculture case studies. In 2021—
Projects from around the world. FAO. https://doi.org/10.4060/cb5359en
Food and Agriculture Organization. OCDE. (2017). Adopter une approche territoriale dans les
politiques de sécurité alimentaire et nutritionnelle. ISBN: 978-92-64-27243-9.
Fuglie, K., Gautam, M., Goyal, A., & Maloney, W. F. (2020). Harvesting prosperity: Technology
and productivity growth in agriculture. World Bank.
Han, J., Kamber, M., & Pei, J. (2011). Data mining: Concepts and techniques (3rd ed.). Morgan
Kaufmann Publishers.
Harbouze, R., Pellissier, J.-P., Rolland, J.-P., & Khechimi, W. (2019). Rapport de synthèse sur
l’agriculture au Maroc. Rapport de Recherche CIHEAM-IAMM, 2019, 104. ‌hal-02137637.
Heady, D., & Fan, S. (2011). Reflections on the global food crisis, IFPRI [Research monograph].
Lionboui, H., Boudhar, A., Lebrini, Y., Htitiou, A., Elame, F., Hadria, R., & Benabdelouahab,
T. (2022). Digitalization and agricultural development: Evidence from Morocco. In M. Behnassi,
M. B. Baig, M. T. Sraïri, A. A. Alsheikh, & A. W. A. Abu Risheh (Eds.), Food security and
climate-smart food systems. Springer. https://doi.org/10.1007/978-­3-­030-­92738-­7_16
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In Climate change:
The social and scientific construct by Suhaib A. Bandh. Springer Nature Publishing. https://
link.springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
MAPMDREF. (2017). Contrats programmes pour le développement des filières de production,
2015; Royaume du Maroc, Stratégie nationale de développement durable, 2030 (octobre).
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
O’Sullivan, M. (2000). The innovative enterprise and corporate governance. Cambridge Journal of
Economics, 24(4), 393–416. https://doi.org/10.1093/cje/24.4.393
OCDE-FAO. (2016). Perspectives agricoles de l’OCDE et de la FAO, 2016–2025. Food and
Agriculture Organization.
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Pray, C. E., & Nagarajan, L. (2014). The transformation of the Indian agricultural input indus-
try: Has it increased agricultural R&D? Agricultural Economics, 45(S1), 145–156. https://doi.
org/10.1111/agec.12138
Rijswijk, K., Klerkx, L., & Turner, J. A. (2019). Digitalisation in the New Zealand agricultural
knowledge and innovation system: Initial understandings and emerging organisational responses
to digital agriculture. NJAS, 90–91(1), 1–14. https://doi.org/10.1016/j.njas.2019.100313
12  Technological and Managerial Innovation in Agriculture to Ensure Food Security… 219

Rockström, J., Steffen, W., Noone, K., Persson, Å., Chapin, F. S. I., III, Lambin, E., Lenton, T. M.,
Scheffer, M., Folke, C., Schellnhuber, H. J., Nykvist, B., de Wit, C. A., Hughes, T., van der
Leeuw, S., Rodhe, H., Sörlin, S., Snyder, P. K., Costanza, R., Svedin, U., et al. (2009). Planetary
boundaries: Exploring the safe operating space for humanity. Ecology and Society, 14(2), 32.
https://doi.org/10.5751/ES-­03180-­140232
Royer, A., de Marcellis-Warin, N., Peignier, I., Warin, T., Panot, M., & Mondin, C. (2020). Les
enjeux du numérique dans le secteur agricole, Défis et opportunités. CIRANO, Rapport de
Projet, 2020RP–2012. ISSN 1499-8629.
Şimşit, Z. T., Vayvay, Ö., & Öztürk, Ö. (2014). An outline of innovation management process:
Building a framework for managers to implement innovation. Procedia – Social and Behavioral
Sciences, 150, 690–699. https://doi.org/10.1016/j.sbspro.2014.09.021
Ton, G., Vellema, W., Desiere, S., Weituschat, S. & D’Haese, M. (2018). Contract farming
for improvingsmallholder incomes: what can we learn from effectiveness studies? World
Development, 104, 46–64. https://doi.org/10.1016/j.worlddev.2017.11.015
Touzard, J. M. (2018). L’innovation agricole et agroalimentaire auxxie siècle: Maintien, efface-
ment ou renouvellement de ses spécificités? Éditions Quæ. http://books.openedition.org/
quae/25291. ISBN: 9782759230266 (pp. 39–55).
United Nations. (2017). Garantir la sécurité alimentaire d’ici à, 2030 : le rôle de la science, de la
technologie et de l’innovation. Conseil économique et social, E/CN.16/2017/3.
UNTCAD. (2017). The role of science, technology and innovation in ensuring food security by
2030. United Nations conference on trade and development.
World Bank. (2016). Who are the poor in the developing world? [Policy research paper]. Poverty
and Equity Global Practice Group.
Yatribi, T. (2020). Innovations technologiques, entrepreneuriat et agriculture: Enjeux, Atouts et
contraintes pour l’agriculture marocaine. Revue de l’Entrepreneuriat et de l’innovation, III(9),
A1V3N9A2020.
Chapter 13
Oyster Farming and a Worldwide
Referendum on Global Carbon
Fee-and-Dividend

Keith McNeill

Abstract  The author managed an oyster farm in Canada and did research on oyster
culture in Sierra Leone in the early 1970s. He uses those experiences to argue for
global carbon fee-and-dividend as a solution to climate change today.
The world is a village with one big, collectively owned oyster bed, he says. The
villagers rely on there being enough oysters to eat. However, some take more oys-
ters than they should and the oyster bed is in danger of collapse.
The solution to this “tragedy of the commons” is to manage the oyster bed as a
common pool resource.
Specifically, the author recommends charging everyone a fee based on the ton-
nage of oysters they harvest. That fee would gradually be raised high enough so that
the harvest never again gets above sustainable levels.
The money collected would go into a common pot and be distributed to everyone
as equal dividends. Because the cost of oysters would be higher than before, villag-
ers would tend to spend less money on oysters and more on alternative sources of
protein.
How do we decide? As people have decided all around the world for many thou-
sands of years – by gathering everyone together and holding a village meeting.

1 Part One: Oyster Farming

I am a bit different. I suppose I have a certain type of intelligence but I am also


socially awkward, physically clumsy and tend to speak in a monotone.
Growing up, while I think most of my classmates and teachers liked and respected
me, there were always a few in school and sometimes even teachers who seemed to
enjoy bullying me.
I thought that things would change when I grew up but that proved not to be
the case.

K. McNeill (*)
Clearwater, B.C., Canada

© The Author(s), under exclusive license to Springer Nature 221


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_13
222 K. McNeill

A few weeks before I graduated from university in 1971, my father purchased an


oyster farm near Fanny Bay on Vancouver Island in British Columbia, Canada.
I still am not clear about why. My career path after graduation had been to
become a medical doctor, not an oyster farmer.
A lot of it had to do with wanting to help my older brother, Johnny. He was
9 years older than I was and, I recognize now, suffering from chronic depression.
He also suffered from chronic pain in his neck, the result of a miscarried parathy-
roid operation several years earlier.
The oyster farm had about 25 hectares of oyster lease, which were areas of beach
granted by the provincial government on, as I recall, a 20-year basis. We paid for
those on a per hectare basis.
We also made use of permits to harvest oysters from public beaches. We paid for
the permits based on tonnes of oysters harvested.
The people we bought the oyster farm from had just one regular customer, a fish
wholesaler in Vancouver. Two weeks after we took over the customer advised us
that he was going to buy all his oysters from another grower.
That was a problem. BC Oyster Board, unlike most marketing boards in Canada,
only set the price but did not provide a quota. That meant that we could not cut our
prices to get customers, only sell on the basis of reputation  – but we had no
reputation.
There was also a BC Shellfish Marketing Cooperative. I worked with others to
make it more active but it was a long while before we got consistent orders through it.
The first year was tough. I was supposed to be helping my older brother but
working in the cold and wet, plus lifting heavy loads did not help the chronic pain
in his neck. That meant that most times I ended up living and working on my own.
We sold a few oysters through a coffee shop that was on the property plus we got
the occasional order from the cooperative and elsewhere. In order to save money for
medical school I did as much of the work as I could by myself.
One advantage I have is my ability to concentrate. I was able to live, sleep and
breathe oysters 24/7. I often put in 14- and 16-hour days, went home to an empty
house and made my own supper and then did the same thing the next day and the
next – for weeks on end.
Eventually, my perseverance paid off. By the third year I had five people working
for me full time and several more part-time.
“The Limits to Growth” came out in 1972 and I was dumbstruck when I read it.
Unless we establish a state of global equilibrium, human civilization faces a sud-
den and uncontrollable collapse, most likely sometime in the first half of the twenty-­
first century.
The authors identified two positive feedback loops with negative consequences
for us: population growth and industrial growth.
We have some ideas about how to control population. People in communities
with restricted resources have been doing that for centuries if not millennia.
However, how would we make it so that investment into industrial capital equals
depreciation? It might be easy enough to do in a computer program but how would
we do it in real life?
13  Oyster Farming and a Worldwide Referendum on Global Carbon Fee-and-Dividend 223

I came across Kenneth Boulding’s essay, “The Economics of the Coming


Spaceship Earth”. In it, he compares planet Earth to a giant spaceship travelling
through the galaxy, with a finite amount of resources in its storerooms and a finite
capacity to absorb garbage and other pollutants.
He suggested that one way to control the immediate problem of pollution was
through resource taxes. As an oyster farmer, that made perfect sense to me. People
would pay per tonne for what they used, similar to how we paid per tonne for the
oysters we picked from public beaches. The more something costs, the less likely
people are to waste it.
Such as system would favour the wealthy, at least in the short term. How to make
it politically acceptable and fair were questions I did not yet have the answers for.
Things were going well at the oyster farm but I was still determined to go to
medical school.
I wrote the Medical College Admissions Test and scored in the 92nd percentile
for verbal ability, 94th percentile in general information, 71st percentile in science
and 25th percentile in quantitative ability.
I did a quick math review and a few weeks later wrote the Graduate Record
Exam, scoring in the 95th percentile in verbal ability and 82nd percentile in quanti-
tative ability. I also scored in the 97th percentile in biology.
I applied to several medical schools and did an interview for the one at University
of British Columbia. My application got as far as the final committee when a uni-
versity official noted that my degree was from Notre Dame University of Nelson, a
small college in the BC Interior. He said that they automatically deduct 20% from
NDU grades. That ended my chances.
I received the phone call telling me committee decision one morning sometime
in the spring of 1973. The afternoon of that same day I received another phone call,
this time from a person representing the International Development Research Centre
(IDRC), a Canadian government agency that funds overseas research.
The IDRC representative asked me if I wanted to be technical advisor on an oys-
ter culture research project in Sierra Leone, West Africa. After learning more about
the proposal, I agreed to go.
Essentially, I was lied to by people I should have been able to trust. But more
about that later.

2 Part Two: Sierra Leone

I was in Sierra Leone for pretty well all of 1974 and 1975.
Before leaving for Sierra Leone I had discussed with several of my employees
how we might convert the oyster farm into a workers’ cooperative when I got back.
The basic idea of a cooperative is that every shareholder has just one vote, no matter
how many shares he or she might hold. The setup would be similar to how fishermen
in former times would each buy shares in a boat and then elect one of their own to
be captain.
224 K. McNeill

Just 2 or 3 weeks before I was to leave for Africa, they told me they would not
accept my offer. I worked too hard and so would end up controlling everything, they
said. One of them said I was “crazy” to put in the number of hours that I did.
Usually, the problem with cooperatives is freeriders  – people who do not do
enough work. In this case, my employees saw the problem as being that I did too
much work.
So, I left for West Africa without a clear understanding on what the oyster farm’s
future would be.
After about 3  months the people I had left in charge quit. Their pay was not
adequate for the 14–16-hour days they sometimes worked, they said. Johnny took
over. He had to hire people to do the work he could not handle.
That same month I met an Englishman who had many decades of experience
working overseas. I told him that I was technical advisor on an oyster culture proj-
ect, that I was not in charge of the project and that I was the only Canadian working
on the project in-country.
“You must have powerful enemies at IDRC,” was his comment.
“I barely know anyone there,” I replied.
“Nevertheless, you have been set up as a scapegoat,” he said. “You will never get
any work out of the Sierra Leoneans. The project will fail and you will get blamed.
You will never work for IDRC again.”
I was shocked by what he said and hoped that I would never be that cynical. I
wish I had listened to him.
Also in March 1974, David Gordon Wilson, a MIT engineering professor, wrote
a letter to the editor of the Christian Science Monitor advocating that a tax or fee be
put on petroleum products, with the money collected being given as equal payments
to every American adult.
I had a subscription to CSM at the time, even though I was in Sierra Leone, and
I remember reading the letter.
What he was proposing was essentially the same as what is know today as carbon
fee-and-dividend, promoted by Citizens Climate Lobby, Dr. James Hansen and oth-
ers as a method to control climate change.
Wilson was writing in response not to climate change but to gasoline shortages
that were happening in the United States and other nations as a result of an Arab oil
embargo caused by the 1973 Mid East War.
I saw that his suggestion solved the issues with Boulding’s resource taxes and so
could be used to control global warming – except for one problem – his proposal
only involved the United States.
How would resource taxes at the national level work in a place like Sierra Leone?
There were rumours of sizeable oil reserves offshore. Would it be reasonable to
expect the Sierra Leoneans to impose a fossil fuel fee on themselves and forego the
income they would get from selling that oil?
I had a Eureka moment – global carbon fee-and-dividend – charge a fee on fossil
fuels, similar to a carbon tax, and return all the money collected as equal dividends
to everyone, similar to a universal basic income – and do it on a global basis, mean-
ing all the money goes into one pot and is distributed equally worldwide.
13  Oyster Farming and a Worldwide Referendum on Global Carbon Fee-and-Dividend 225

Even though my contract was for 2 years, my plan had been to just stay in Sierra
Leone for one. I had told the people from IDRC this from the beginning but, when
the end of my first year approached, they invited me back to Canada for consulta-
tions. Christmas was coming and my contract included several weeks of paid leave
plus a trip home each year, they pointed out.
I went back to Canada and they poured on the charm. Everything I had done so
far was wonderful and beyond expectations. An oyster culture project was starting
in Sabah. Before I had left, they had hinted to me that perhaps I could get a second
contract there. Now it was presented as a virtual certainty.
I was told that IDRC provided professional development scholarships to its con-
tractors, which meant that I could pursue a master’s or doctorate in marine biology
and the agency would pay the way. They even gave me a raise in my salary.
The oyster farm was practically dead and it did not seem that another year would
make it worse. So, I accepted and went back to Sierra Leone for a second year.
A strange thing happened a few months into my second year. One of the people
from IDRC came for a visit. He set up an elaborate joke at my expense in front of
the project’s crew. The clear intent was to humiliate me in front of them. He had a
good laugh but the crew simply sat there in silence with “no expression” expres-
sions on their faces.
I recalled the Englishman’s warning. I realized that this person, who had been so
charming and who now held my future in his hands, in fact disliked me. I knew with
certainty that he intended to ruin my life and very likely would succeed.
I really liked the Sierra Leoneans. I found them to be cheerful, friendly and hard-
working people.
At the time, Sierra Leone ranked near the bottom of the list on several quality of
life indices. What made it different from those near the top?
I learned the importance of worldview or paradigm.
The Sierra Leoneans would not put up with petty thievery. I remember watching
a policeman frog-marching a thief towards the local jail. The thief had been caught
taking something at a nearby market. A small mob followed along. They shouted at
the thief, threw garbage at him and tried to kick or punch him. It was clear that, if
the policeman had not been there to protect him, the thief would have been given a
severe beating.
The Sierra Leoneans did not tolerate petty thievery but they put up with large-­
scale bribery and corruption.
I suspect that part of the reason was the slave trade. For hundreds of years, if you
wanted to become rich and powerful in West Africa, you sold your neighbour to the
white man as a slave. Why would you want to do that? Because if you did not, there
was a good chance your neighbour would do it to you.
Some of the same families that grew rich during the slave trade still rule in Sierra
Leone today. That elite’s worldview has become the worldview of the entire society,
even though it has negative consequences for nearly everyone. (Incidentally, I
understand things have improved somewhat but that corruption is still a problem.)
At the end of my second year I went straight to IDRC headquarters in Ottawa.
The innovations I had done in Sierra Leone to get the project off the ground, which
226 K. McNeill

had been the subject of so much praise at the end of my first year, now were con-
demned as unauthorized deviations from my instructions.
As for the Sabah project, I was told that the Sabahan staff were fully qualified
and there would be no need for my services.
IDRC did not provide professional development scholarships to its employees
and contractors, I was told. The only IDRC scholarships were postdoctoral fellow-
ships to people from developing countries.
I was made a scapegoat, just as the Englishman had predicted.
The irony was that the Sierra Leoneans and I had worked hard and we had come
as close as anyone in the world to figuring out how to grow tropical oysters on a
commercial basis for food.
I went back to Fanny Bay. Johnny had tried but it had been too much for him. The
number of customers had shrunk to just one chain of fish stores, The Big Crab.
A few weeks later even that ended when The Big Crab went bankrupt.
After more than 5 years of hard work I found myself back where I had started,
only worse. Back then, the oyster farm had no customers and no reputation. Now, it
had no customers and a reputation of not filling orders. I started doing the rounds of
the fish wholesalers but without success.
Other bad things happened as well, things that I will not go into.
One evening as I was trying to sleep, I heard people talking just outside my door.
I got up but it was just the noise of traffic on the nearby highway. I went back to bed
and, again, I distinctly heard voices. I got up a second time and again found that was
just traffic noise. This went on for several hours.
Hearing voices that were not there frightened me. I believe in pushing myself but
also am aware of the need not to push myself so far that something breaks. It seemed
to me that I was approaching that point with my mental condition. What to do?
Future studies in medicine or marine biology seemed out of the question. I could
not very well fail to mention my time in Sierra Leone in my application but, if I did,
those looking at my application could hardly fail to wonder why I did not include
anyone from IDRC as references.
I left the oyster farm for Johnny to run. He did so on a part-time basis until he
died of carbon monoxide poisoning in 1978.
My father sold the oyster farm to a partnership that included the former employ-
ees who had refused my workers’ cooperative proposal several years earlier. They
changed its name to Fanny Bay Oyster Company and in turn sold it a few years later
to Taylor Shellfish, a large oyster company from Washington State. The business is
still in operation and, when I visited it a few years ago, appeared to be thriving.

3 Part Three: Clearwater

After living in several different communities, working at different jobs and looking
at various business opportunities, in 1981 I found myself in Clearwater, BC with
something like $50 in my pocket.
13  Oyster Farming and a Worldwide Referendum on Global Carbon Fee-and-Dividend 227

I started work as a typesetter with the Clearwater Times newspaper. Soon I was
working as a reporter and photographer as well.
I rented and then bought a small log cabin and began spending a lot of time
reading.
Changing a group of people’s paradigm or worldview is difficult, as I saw in
Sierra Leone. Hard knocks had changed mine somewhat. What would it take to
change the whole world’s?
Paradigms or worldviews can appear to change suddenly, when the situation
makes it obvious that they are no longer adequate. However, getting them to change
in the right direction requires years of preparation and gentle nudging.
Successfully implementing global carbon fee-and-dividend would require a
stronger and more democratic United Nations. The UN and its predecessor, the
League of Nations, were born painfully out of the World War II and World War I,
respectively. We cannot afford a World War III. What examples do we have in his-
tory of major political reforms being achieved peacefully?
One is Hiawatha, also known as Ayenwathaaa or Aiionwatha. He and his associ-
ate, Dekanawidah, travelled among the tribes in what is now the northeastern United
States and eastern Canada to unite them into the Iroquois Confederacy or
Haudenosaunee. When exactly this happened is not clear but it happened long
before the arrival of Europeans.
The confederation they co-founded later served as a model of federal governance
for the United States and for Canada.
Another is Mahatma Gandhi, who used nonviolent techniques to free India from
British rule.
Possibly Gandhi’s greatest triumph was the Salt March of 1930. He and several
dozen followers walked 385 km from an ashram near Ahmedabad to the sea. There
he made salt and sold it without paying the salt tax. He was promptly arrested.
Thousands of others across India did the same and also were arrested. Gandhi’s
argument was everyone needed salt but the poor need it more than anyone. The salt
tax was a tax on the poor man’s sweat, he said.
As a tactic it was brilliant. Rather than directly attack British rule, he focussed on
just one aspect of it that was morally indefensible but that the regime could not do
without.
The Chartists of England took their name from the People’s Charter of 1838.
They presented petitions to parliament with millions of signatures calling for,
among other things, universal manhood suffrage. Although the movement was
mostly peaceful, several of leaders were imprisoned or transported to Australia.
Eventually, however, nearly all of their demands were met.
In Switzerland, if 100,000 citizens sign a petition calling for a proposition, the
government is obliged to hold a referendum on the question. If the referendum
passes, the proposition becomes legally binding.
Where did Swiss direct democracy come from? Two of the 26 cantons, Appenzell
Innerhoden and Glarus, still are governed by open air assemblies or Landesgemeinde
held annually. Similar open air assemblies are held at the village or communal level
in some cantons.
228 K. McNeill

In early 1987 the young son of a relative cried himself to sleep after watching a
documentary about climate change on TV.
I knew that an essential part of the solution to climate change had to global car-
bon fee-and-dividend but I was doing nothing to promote it. Where to begin?
I resolved to walk around the world to India and back as a pilgrimage to honour
Mahatma Gandhi. Perhaps that would stir interest in using his approach to deal with
global issues.
I left Clearwater in June 1987. During the next 6 months of that year and 2 months
in 1989, I walked as far as Toronto, a distance of about 4000 km. Foot problems
forced me to go back to work at the Times but I was not there long.
New owners took over in 1991 and they did not like my environmental views. I
was laid off. I did various odd jobs for 5 years, including teaching adult education
at a local prison camp and being first aid man on a forest firefighting crew.
In 1996, I returned to the Times, this time as editor. Black Press, a BC-based
newspaper chain, had bought it out.
I retired from the Times in 2018. During my time as editor the newspaper won
best all-round newspaper in its circulation class from Canadian Community
Newspapers Association two times and the equivalent award from BC and Yukon
Community Newspapers Association three times.
In 2015, I started an online petition calling for carbon fee-and-dividend in
Canada. It went viral and within about 3 or 4 weeks had collected 28,000 names.
To promote the petition, fellow environmentalist Jean Nelson and I cycled from
Toronto to Ottawa for a Citizens Climate Lobby-Canada conference. Along the way
we did interviews with several community newspapers to talk about the petition and
our bike trip.
The Canadian federal government passed its Greenhouse Gas Pollution Pricing
Act in 2018, 3 years after our bike ride and petition. The act imposed a carbon tax
in those provinces that did not have a carbon price of their own. With 90% of the
money being returned to households, it is essentially carbon fee-and-dividend.
Catherine McKenna, Canada’s environment minister at the time, played a major
role in stick-handling the bill’s passage. Her riding of Ottawa Centre was where
Jean and I ended our ride. Maybe she read about the 28,000 names on our petition
in one of our newspaper interviews or maybe one of her staff did. There no doubt
were many factors involved, with the biggest credit going to CCL-Canada and its
director, Cathy Orlando. Still, it is nice to think that our effort might have been the
nudge that made carbon fee-and-dividend happen in Canada. Who knows?
1FannyBayKeithMcKenKeithman
Author Keith McNeill (r) takes a break with neighbour Ken Keithman as they pick oysters from a
public beach near Fanny Bay, BC. The government charged a fee per tonne for the oysters. A simi-
lar approach should be used to price carbon dioxide globally, says the author. (Photo courtesy of
Keith McNeill)

2SL-AbuSawyerAndersonVidal (L-r)
Fisheries officer and project leader Abu Kamara, project worker Sawyer and project mechanic
Anderson watch as Vidal Forde, a faculty member at Fourah Bay College, collects a water sample
230 K. McNeill

to test for salinity in Number Two River near Freetown in 1974. The project later acquired an
electronic device to measure salinity. (Photo courtesy of Keith McNeill)

2SL-DixonBulletHole
Fisheries officer Dixon points to a bullet hole in an oil barrel that had been supporting one of the
project’s oyster rafts near Jui. A member of the ISU, a paramilitary police force, had asked for a
bribe to ensure nothing unfortunate happened to the rafts. When a bribe was not forthcoming, bul-
let holes appeared in the rafts. The ISU member was transferred to a remote post upcountry. (Photo
courtesy of Keith McNeill)
2SL-KeithOysterStringsNumber2River
The author, Keith McNeill, stands next to strings of oyster shells in Number Two River village in
1974. The strings were hung from rafts where they would collect oyster spawn. The tiny oysters
then grew quickly to market size. At the time a quiet fishing community with a beautiful beach, the
Number Two River is now an important tourist destination. (Photo courtesy of Keith McNeill)

3Ottawa-ElizMayKeithJean (L-r)
MP Elizabeth May, then the leader of the Green Party of Canada, receives a USB stick from Keith
McNeill and fellow environmentalist Jean Nelson containing a petition with 28,000 names calling
for carbon fee-and-dividend in Canada in 2015. Three years later Canada’s Liberal government
brought in a carbon tax with rebates scheme that is very similar to carbon fee-and-dividend. (Photo
courtesy of Keith McNeill)
232 K. McNeill

References and Further Reading

Boulding, K. (1966). The economics of the coming spaceship Earth. In H.  Jarrett (Ed.),
Environmental quality in a growing economy (pp.  3–14). Resources for the Future/Johns
Hopkins University Press.
Guha, R. (2018). Gandhi: The years that changed the world. Alfred A. Knopf.
Hsu, S.-L. (2011). The case for a carbon tax. Island Press.
Kamara, A. B., & McNeill, K. (1976). Oyster culture project: Preliminary oyster culture experi-
ments in Sierra Leone. Fisheries division occasional Paper no. 1, Ministry of Agriculture and
Natural Resources, Freetown, Sierra Leone.
Meadows, D. H., et al. (1972). The limits to growth: A report for the Club of Rome’s project on the
predicament of mankind. Signet book from New American Library.
Wilson, D.  G. (1974). Here’s a new alternative to gas rationing. Christian Science Monitor,
March, 15.
Ostrom, E. (1990). Governing the commons: The evolution of institutions for collective action.
Cambridge University Press.

Websites

Citizens Climate Lobby. https://citizensclimatelobby.org/


Citizens Climate Lobby-Canada. https://canada.citizensclimatelobby.org/
Citizens Climate International. https://citizensclimate.earth/
Democracy Without Borders. https://www.democracywithoutborders.org/
World Basic Income. https://www.worldbasicincome.org.uk/
Chapter 14
Climate Change Impact Modeling
on Citrus Yield

Fouad Elame, Youssef Chebli, Hallam Jamal, and Lionboui Hayat

Abstract  According to many scientific institutions, rising temperature and precipi-


tation decrease, in addition to the increase of extreme weather events, will have a
significant effect on agriculture in the Mediterranean region by the end of this
century.
The citrus sector in Morocco plays a significant socio-economic role. It is the
most important source of income creating more than 16 million working days annu-
ally and ensures an average production of two million Tons per year (2,275,400 T in
2018). The development of this crop is achieved at the expense of a resource that
will become scarce in the long term.
Indeed, water is an important production factor in agriculture. Its availability, or
its shortage, can lead to a successful harvest, or a decrease in yields, or even its
failure. According to the FAO, irrigation basically doubles agricultural yields and
the number of crops grown in a year. However, climate change will have a drastic
impact on irrigated crops over the coming years, citrus among others. In order to
support the agricultural sector towards better risk management, this chapter
describes climate variation, analyzes the value chain of citrus in Morocco, and per-
forms a statistical analysis of the impact of climate change on the evolution of citrus
areas and yields.

Keywords  Climate change · Impact · Agriculture · Citrus · Yield

F. Elame (*) · H. Jamal


INRA, Regional Center of Agronomic Research of Agadir, Agadir, Morocco
e-mail: fouad.elame@inra.ma
Y. Chebli
INRA, Regional Center of Agronomic Research of Tangier, Tangier, Morocco
L. Hayat
INRA, Regional Center of Agronomic Research of Rabat, Rabat, Morocco

© The Author(s), under exclusive license to Springer Nature 233


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_14
234 F. Elame et al.

1 Introduction

The problem of climate change is undeniably the phenomenon of the century. The
last three decades have been the warmest than the other decades since 1850, of
which the period 2001–2010 was marked by the most intense heat waves (GIEC,
2014). This change is expressed by the variability of rainfall patterns, the increase
in temperature, and the risk of extreme events such as droughts, floods, and cyclones.
The North African region is a sensitive area to climate change. According to the
World Bank (2010), dust storms and heat wave risks are likely to make this region
absolutely unbearable; in addition to the decrease in rainfall that has already been
observed, with a rise in temperature of 2 °C, between 20% and 40% of rainfall in
North Africa should disappear.
Agriculture is expected to be one of the most impacted economic activities by
climate change, as variations in temperature and rainfall generally result in reduced
yields of many crops (Gommes et al., 2009). Sustainable agricultural land use is
crucial to global food security (Lionboui et al., 2022). Indeed, the agricultural pro-
ductivity growth trend in the last decades has been enormous in many ways, which
has helped to alleviate poverty and food insecurity in many areas. This was mainly
due to improved production systems and large investments (Tyner & Ouraich,
2014). Nevertheless, climate change threatens to reinforce the existing challenges
facing agriculture. The world population is estimated to reach 9 billion before 2050,
taking into account the accelerating demand for food; the FAO predicts that the food
production will need an increase of 70% to meet the world’s population demand
(FAO, 2011; Malla et  al., 2022; Bandh et  al., 2021, 2022, 2023; Mushtaq et  al.,
2020; Parray et al., 2022; Bandh, 2022a, b).
As a country located in northern Africa, Morocco is distinguished by four types
of climates: humid, semi-humid, semiarid, and arid. Nevertheless, the latter two
tend to prevail in several regions to the detriment of humid and subhumid climate
areas which are experiencing a remarkable regression (HCP, 2014). The forecasts of
the Intergovernmental Panel on Climate Change (IPCC) indicate an increase in
average temperatures in the summer period of around 2.2–5.1 °C as well as an aver-
age decrease in precipitation of −4% to −27% by the end of this century (GIEC,
2007). The agricultural sector is undoubtedly impacted by climate change, due
mainly to water stress and the accentuation of extreme waves of drought and floods.
The country’s economy is generally affected by these effects, due to the fact that
agriculture contributes with 19% of the gross domestic product (GDP) and 80% of
direct rural employment, without neglecting the issue of food security
(MADRPM, 2017).
The citrus sector in Morocco plays a very important socio-economic role. The
citrus sector occupies an area of ​​127,100 ha and estimates an average production of
more than 2.4 million tons per year (HCP, 2019). According to the ministry of agri-
culture, this sector contributes to improving farmers’ incomes as well as being an
important source of job creation, around 15.9 million working days in 2019, most
14  Climate Change Impact Modeling on Citrus Yield 235

of it in orchards and the rest in the processing industry, packaging, and other activi-
ties related to the value chain. At the economic level, the citrus industry is a signifi-
cant source of foreign currency around 4.8 billion Dirhams (Dh), recorded in 2019,
coming from exports that are nearly 644.000 tons to various countries mainly the
European Union and Russia.
Citrus fruits have a major socio-economic weight in Morocco, particularly in the
Souss-Massa, the area of ​​which is the most dominant in Morocco. However, it
remains very vulnerable to climatic risks. In a region where the arid and semiarid
climate is the most dominant, and the scarcity of water resources is the most remark-
able, the yield of citrus fruits is threatened by climate change. Therefore, assessing
the impacts of climate change on citrus production in the Souss-Massa region seems
to be relevant.
The objective of this chapter is to assess the main determinants of citrus produc-
tion in the Souss-Massa region and the factors influencing its productivity, through
an econometric approach. This modeling approach makes it possible to analyze the
annual variation in citrus production according to the determining factors, as well
as to identify the importance of each determining factor and its level of influence.
This work also aims to predict the main impacts of climate change on the cit-
rus sector.

2 Climate Change Impacts and Agriculture

The concept of climate change entered public consciousness as early as the 1990s
(Beniston, 2009). The austerity of the impacts of this change is increasingly sig-
nificant on the international level, which has prompted governments to launch
debates and actions to address this problem. Recently, human beings are becoming
concerned about their survival resources that the earth and the environment pro-
vide for them, such as water, air, and food. Therefore, climate change brings into
question deep ethical issues, among them, the ability of the current generation to
leave an acceptable environmental legacy to future generations. Indeed, agriculture
is one of the human activities that occupies an important place knowing that its
impact on natural resources could be significant; in addition to playing a signifi-
cant role in the development of territories, it meets one of the most basic needs
which is food.
Rising temperature and declining precipitation, in addition to the increase in
extreme weather events, will have a significant effect on agriculture in the
Mediterranean region by the end of this century. On a global scale, agriculture,
deforestation, and other land uses are responsible for around 23% of greenhouse gas
emissions. Facing these facts, ensuring food security for the growing populations
becomes a worldwide challenge (IPCC, 2019). Agriculture will have to adapt to
new contexts and tackle most of these changes.
236 F. Elame et al.

2.1 Climate Change Is a Fact

Climate change is defined as a variation in the mean state of the climate or its vari-
ability over a long and continuous period that reaches decades or more (GIEC,
2014). The origins of this climate change are diverse, namely, those due to internal
processes in the climate system or to external natural or anthropogenic forcings.
Human activity linked to industry, agriculture, energy production, and transport
has led to the emission into the atmosphere of significant quantities of greenhouse
gases (GHGs) since the beginning of the industrial era of the nineteenth century.
Nowadays, the global climate is reacting to this increase in GHGs. The first IPCC
report concluded that global average temperatures could increase by 1.5–5.8 °C by
the end of the twenty-first century in response to increased radiative atmosphere
forcing (GIEC, 2007). Although this seems to be a negligible warming in compari-
son to the diurnal and seasonal amplitude of temperatures, it must be remembered
here that the increase in the predicted global average temperature has been unprec-
edented for at least 10,000 years. It is not only the magnitude but also the rate of
warming that worries the scientific community, mainly because of the vulnerability
of environmental and socio-economic systems to the speed of these changes (Harris
et al., 2017).
There are many principal reasons that explain anthropogenic influence on the
climate and the natural ecosystem. Most are linked to complex interconnection
between economic, political, and cultural factors. However, economic development
and demographic growth are the two major factors explaining most of the pressures
exerted by humans on our ecosystem.
The impact of human activities on the emission of greenhouse gases from fossil
fuels, industrial processes, and agriculture is significant. Indeed, the effects of
GHGs on socio-economic and ecological sustainability have already been seen and
will continue to grow at the regional as well as global level in the coming years.
Indeed, agriculture has contributed in the variation of our climate, for example, the
agricultural land management mainly manifested by the conversion of forests to
agricultural land and by the emission of GHGs. In addition to the emission of CO2
by the introduction of machinery to this sector, livestock and rice fields also have a
deep influence on the concentration of GHGs by significant emissions of methane
(CH4) (World Bank, 2010).

2.2 Impact of Climate Change on Agriculture

The variability of crop yields has been deeply studied in the literature in relation to
various production factors. Agriculture depends on water as the most important pro-
duction factor. Biomass production is closely linked to water requirement, and live-
stock depend on water for drinking. Therefore, water is an essential climatic factor;
it affects or determines the growth of plants. Its availability, or its shortage, can
mean a successful harvest, or a decrease in yields, or even its failure. According to
14  Climate Change Impact Modeling on Citrus Yield 237

the FAO (2011), irrigation doubles crop yields and allows growing more than one
crop in a year. Without irrigation, land abandonment, possible relocation of agricul-
tural production, and serious economic problems will be difficult phenomena to pin
down. Hence, to protect agricultural production in some areas requires a scientific
knowledge of climate indicators and better irrigation techniques and practices that
promote water saving (Elame et al., 2017).
The effects of climate change will be felt in many ways at the water level. The
water cycle will be globally affected. If the world must become wetter because
global warming will accelerate the hydrological cycle, the increase in evaporation
will also increase the prevalence of drought conditions. In most regions, precipita-
tion will be more intense and more variable and will often be interposed with longer
dry periods. This development will have widespread effects on human activities and
natural systems. It is possible that changes will occur so rapidly and unpredictably
that traditional agricultural and water management practices will cease to produce
results (FAO, 2017; World Bank, 2010).
Climate change also threatens food security, directly through changes in tem-
perature and precipitation patterns, or indirectly through the loss of agricultural land
due to sea level rise, as well as erosion due to wind and water.
By its nature, agriculture is among the sectors most vulnerable to climate changes
and their impacts. Adaptation is an important component of any action to reduce
climate change. Several studies show that without adaptation, climate change is
generally problematic for the agricultural sector and regional economies; delayed
adaptation actions would be more costly and sometimes fatal.

2.3 Citrus Growing and Environmental Challenges


in the Souss-Massa Region
2.3.1 The Peculiarities of Citrus Fruits

As part of the Morocco Green Plan, a program contract was signed in 2008 between
the government and the private sector for the citrus sector. This initiative has pro-
moted the development of the sector thanks to the efforts undertaken by the profes-
sionals and the incentives granted by the government and relating essentially to the
encouragement of investment intended for the improvement of the productive
upstream, the modernization of the production systems, and the promotion of exports.
This has supported the adaptation of the citrus industry to the international market
and had offered good-quality fruits to domestic and foreign consumers (HCP, 2019)
(Fig. 14.1).
In Morocco, there are six well-known citrus-growing regions (Souss, Marrakech,
Tadla, Atlantic Coast from Larache to Azemmour, Gharb, and Berkane). With a cur-
rent area of 128,000 Ha and an average production of more than two million T per
year, the citrus sector in Morocco plays an important socio-economic role. The sale
238 F. Elame et al.

2500

2000

1500

1000

500

Production
(1000 tons)

Fig. 14.1  Citrus production in Morocco of 1000 tons (Source: MADRPMDREF, 2019)

40400 40343

40200
40043
40000 39955
39808
39800
39635
39600

39400

39200
2012-2013 2013-2014 2014-2015 2015-2016 2016-2017
Area (Ha )

Fig. 14.2  Total citrus area in the Souss-Massa region

of citrus production concerns three main destinations, namely, fresh domestic con-
sumption, export, and processing.
Citrus cultivation is highly developed in the Souss-Massa area. This ranks first in
terms of citrus area at the national level with 40,343 ha (2016–2017) (Fig. 14.2).
14  Climate Change Impact Modeling on Citrus Yield 239

Table 14.1  Regional citrus production (in tons)


Compagnes 2012/2013 2013/2014 2014/2015 2015/2016 2016/2017
Clémentine 297,200 415,000 275,000 260,000 360,000
Navel 30,600 57,000 38,000 32,000 80,000
Maroc Late 86,700 160,000 125,000 110,000 160,000
Autres 142,800 256,000 162,000 208,000 211,000
Total 689,300 405,100 888,000 600,000 811,000
Source: HCP (2017)

2016/2017 423,173

2015/2016 3,55,511

2014/2015 2,97,408

2013/2014 4,33,813

2012/2013 2,50,362

0 1,00,000 2,00,000 3,00,000 4,00,000 5,00,000


Exports
(Tons)

Fig. 14.3  Evolution of regional citrus exports (in tons)

With a diversified varietal profile to gain in earliness or lateness, the citrus sector
in the region has been able to differentiate its offers by producing mainly Clementine,
Navel, and Maroc Late. The regional production in 2016/2017 was estimated at
811,000 T representing 34% of national production (Table 14.1).
From an economic point of view, citrus exports, which fluctuate around an aver-
age of 423,173 T in 2016/2017, represent 65% of national citrus exports and there-
fore are an important source of foreign currency for the region. The potential for
growth and development of this sector and the important intrinsic assets that the
region has in citrus production meet the specific requirements of several destina-
tions and consumers (European, Russian, and others) (Fig. 14.3).
240 F. Elame et al.

3 Methodological Approach

3.1 Introducing the Study Area

Located in the southwest of Morocco, the Souss-Massa region is both an Atlantic


and continental environment and constitutes the gateway to the desert and covers an
area of around 53,789 km2 with a population of 2,741,224 inhabitants recorded in
2016 (HCP, 2017).
The Souss-Massa region has significant economic potential giving its multiple
and diversified vocations (tourism, fishing, agriculture, industry, mining, etc.).
The climate of the region is predominantly arid but it varies from humid, with
cold winter, in the Western High Atlas, to pre-Saharan, with cool winter, in plain.
Surface water resources are limited and very irregular while the annual precipita-
tions do not exceed 200  mm. Groundwater resources remain overexploited. The
aquifers are experiencing an unprecedented drop recording a deficit exceeding
270 Mm3 (ABHSM, 2018).
The agricultural area represents more than 616,500 ha. Most of these agricultural
lands are located in the Souss-Massa plain, where early crops and citrus are the
main productions (Haddouch et al., 2016). Thus, and because of intensive agricul-
ture and the climate aridity, the largest aquifers of the area are in deficit, which
exposes the region to a risk of marine intrusion due to the proximity to the sea. The
effect of aquifer drawdown on agriculture is very noticeable as some areas have
been abandoned.
It therefore turns out that managing water resources in the Souss-Massa area is
complicated. Taking into account its natural environment and the aridity of its cli-
mate, the Souss-Massa watershed is an area at risk. Economic development induced
by agricultural intensification and the impact of climatic variations indicated by
drought that periodically rages in the region, added to the irregularity of rains and
watercourses, explain the resource scarcity and the recourse to the massive exploita-
tion of groundwater (Fig. 14.4).

3.2 Building the Econometric Model

Multiple linear regression is the most frequently used statistical tool for the study of
multidimensional data and the most dominant analysis method in linear regression.
It is a predictive analysis to explain the relationship between a continuous depen-
dent variable and two or more independent variables. The model resulting from this
type of regression allows the identification of the effect level that the independent
variables have on the dependent variable and can even predict the trends and future
values ​​of the variables (Brown et al., 2020; Nisbet et al., 2018).
Data used in this work are provided by the Regional Office for Agricultural
Development of Souss-Massa and the Souss-Massa Bassin Agency. The data
14  Climate Change Impact Modeling on Citrus Yield 241

Fig. 14.4  Provinces of the Souss-Massa region (DGCL, 2015)

concern citrus production trends at the regional level, average temperatures, and
rainfall as well as the agricultural areas covered by citrus fruits, between 2000
and 2020.
This linear model was developed using a stepwise regression method. It is a sta-
tistical method that uses statistical significance to select the explanatory variables
that build the final statistical model. It involves gradually removing or adding poten-
tial variables explaining the dependent variable and testing for their statistical sig-
nificance (Mark & Goldberg, 2001; Smith, 2018).
We have chosen three explanatory variables of citrus production in the region:
area, temperature, and precipitation. The model will therefore be as follows:

PROD  f  TEMP, PREC, SUP 



With:

PROD: as a continuous quantitative variable representing the production of citrus fruits in the
Souss-Massa region expressed in tons
TEMP: as a continuous quantitative variable representing the average temperature in the
Souss-Massa region expressed in °C
PREC: as a quantitative variable representing the average rainfall in the Souss-Massa region
expressed in mm
SUP: as a quantitative variable representing the area of citrus crops in hectare
242 F. Elame et al.

4 Results and Discussion

4.1 Summary Analysis

Statistical analysis revealed the following results. According to Table 14.2, the cor-
relation coefficient value is 0.719. Indeed, the value R = 71.9% affirms that there is
a significant correlation between the independent variables (PROD) and the depen-
dent variables (TEMP, PREC, SUP). Regarding the coefficient of determination
R2  =  0.509, this value means that the independent variables explain 51% of the
model variability.
After this model summary analysis, we need to conduct the Fisher test “F.” The
critical value of “F” allows to check whether the observed relationship is significant
or due to chance. If less than or equal to the critical threshold, it is significant; oth-
erwise it is due to chance. We generally take a critical threshold of 0.05 (5%). In our
case, 0.019 < 0.05, so it is significant. We conclude that there is a significant rela-
tionship between the dependent variable and the explanatory variables. That is to
say that we have at least one nonzero parameter, so the model is globally significant
(Table 14.3).

4.2 The Regression Equation

The standard error shows the coefficient of variation in the population. It also allows
calculating the value of Student’s T distribution. The larger the Student t-value and
the smaller the p-value, the more the predictor contributes to the model. To carry out
this analysis, we can first observe the “p-value” associated with the Student’s test,
which varies between 0 and 1. Generally, a “p-value” equal to or less than 0.05

Table 14.2  The model summary


Model summarya
Model R R square Adjusted R square Std. error of the estimate
1 0.719b 0.509 0.401 89,958.514
Dependent variable: PROD
a

Predictors: (constant) TEMP, PREC, SUP


b

Table 14.3  Result of the ANOVA


ANOVAa
Model Sum of squares df Mean square F Sig.
Regression 107,599,628,324.077 3 35,866,542,774.692 4.432 0.019b
Residual 129,480,547,490.366 16 8,092,534,218.148
Total 237,080,175,814.443 19
a
Dependent variable: PROD
b
Predictors: (constant) TEMP, PREC, SUP
14  Climate Change Impact Modeling on Citrus Yield 243

Table 14.4  Model results: coefficients


Unstandardized coefficients Standardized coefficients
Model B Std. error Beta t Sig.
(Constant) 1,610,265.833 384,224.356 4.191 0.001
SUP 2.602 4.199 0.120 0.620 0.544
PREC 81.664 153.556 0.101 0.532 0.602
TEMP −53,213.653 16,472.182 −0.644 −3.231 0.005

indicates that the studied variable significantly influences the dependent variable.
We therefore conclude that the temperature has a significant impact on production
since the p-value is low (Table 14.4).
From the table of coefficients, we can conclude that the variable “TEMP” con-
tributes negatively to the explanation of the dependent variable “Prod” with a rela-
tive weight of 64.4%. The independent variable “PREC,” which represents
precipitations, contributes positively to the explanation of the variability of the pro-
duction with 10%. The independent variable “SUP,” which represents the area, con-
tributes positively to the explanation of the production with 12%.
These nonstandardized coefficients allow building the regression equation of our
model and becomes as follows:

Y  2.602 X 1  81.664 X 2  53, 213.65 X 3  1, 610, 265.83 (14.1)



The quality of the model adjustment is average given the existence of other explana-
tory variables not taken into account such as soil treatment, pesticide use, technical
development, and agricultural machinery. We have to mention that the citrus sector
is not a rainfed crop but an irrigated crop which has a very high water requirement.
As mentioned before, the Souss region is the leading national producer of citrus
fruits in Morocco. Nevertheless, this region suffers from an increasingly serious
water deficit exceeding 271 million m3 (Elame et al., 2020) and groundwater pump-
ing continues to exacerbate the deficit causing environmental and economic prob-
lems. Taking into account these findings, we can say that citrus production in the
Souss-Massa area is still on the rise; however, this performance is realized to the
detriment of a resource that is becoming very scarce. In fact, the continued drop in
the water table forces farmers to speed up pace of well deepening, already exceed-
ing 300 meters. In addition to environmental damages, the pressure on groundwater
also has negative economic impacts, of which mainly increased production costs
and saltwater intrusion are the most visible.

5 Conclusions

The Souss-Massa area is well known for its agricultural activities. It is the leading
producer and exporter of citrus at the national level. Since the launch of the
Moroccan green plan in 2008, and because of the significant investments in this sec-
tor, it has experienced tremendous dynamics and performance.
244 F. Elame et al.

To demonstrate the impact of climate change on citrus production in the Souss-­


Massa area, we have computed an econometric linear model. According to our
results, we found that the temperature is a major factor affecting the production.
Indeed, the increase in temperature leads to a drop in production, especially, heat
waves that have been more pronounced these last years, which confirms the influ-
ence of climate change on the sector. Nevertheless, rainfall has relatively no effect
on production since citrus growing is not a rainfed crop. The intensification of citrus
production in the region is achieved to the detriment of groundwater resources. The
region’s water deficit is increasing significantly, while other alternatives must be
considered to overcome this deficit, such as implementing see water desalination,
reuse of wastewater, and encouraging agricultural research and innovation to allevi-
ate climate change impacts.

References

ABHSM. (2018). Agence du Bassin Hydraulique de Souss-Massa. Etude d’élaboration du contrat


de nappe de Souss. Mission 1: Elaboration du dossier préliminaire et organisation de l’atelier
de démarrage. Rapport provisoire.
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S.  A. (2022b). Sustainable agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­a nd-­M icrobial-­D iversity-­A dvances-­a nd-­C hallenges/Bandh-­Parray-­S hameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting cir-cular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Beniston, M. (2009). Changements climatiques et impacts de l’échelle globale à l’échelle locale,
Presses polytechnique et universitaires romandes, 243p.
Brown, S.  D., Tauler, R., & Walczak, B. (2020). Comprehensive chemometrics:
Chemical and biochemical data analysis. Elsevier Science. https://books.google.es/
books?id=heXEDwAAQBAJ. ISBN 9780444641663.
DGCL. (2015). Direction Générale des Collectivités Locales, Ministere de l’Interieur. Région de
Souss-Massa, 2015, 62p.
Elame, F., Lionboui, H., & Choukr-Allah, R. (2017). Water use efficiency and valuation in agri-
culture in the Souss-Massa (pp. 275–283). The Souss-Massa River Basin, Hdb Env Chem 53.
https://doi.org/10.1007/698_2016_75.
Elame, F., Doukkali, R., & Lionboui, H. (2020). Dynamic modeling of climate change impact
on agricultural landsand water resources. In: Leal Filho W., Luetz J., Ayal D. (Eds.),
Handbook of Climate Change Management. Springer, Cham. https://doi.org/10.1007/978-­3-
­030-­22759-­3_41-­1
14  Climate Change Impact Modeling on Citrus Yield 245

Food and Agriculture Organization. (2011). Climate change, water and food. Food and Agriculture
Organization. Water Report no. 36. Food and Agriculture Organization of the United
Nations (FAO).
Food and Agriculture Organization. (2017). Water for sustainable food and agriculture. A report
produced for the G20 Presidency of Germany Food and Agriculture Organization of the United
Nations. ISBN 978-92-5-109977-3.
GIEC. (2007). Bilan des changements climatiques. Rapport de synthèse. Contribution des Groupes
de travail I, II et III au quatrième Rapport d’évaluation du Groupe d’experts intergouvernemen-
tal sur l’évolution du climat. https://www.ipcc.ch/pdf/assessment-­report/ar4/syr/ar4_syr_fr.pdf
GIEC. (2014). Changements climatiques 2014: Incidences, adaptation et vulnérabilité. Suisse. 40 p.
Gommes, R., El Hairech, T., Rosillon, D., & Balaghi, R. (2009). Impact of climate change on agri-
cultural yields in Morocco. World Bank – Morocco study on the impact of climate change on
the agricultural sector. Food and Agriculture Organization of the United Nations (FAO). http://
www.fao.org/nr/climpag/pub/FAO_WorldBank_Study_CC_Morocco_2008.pdf
Haddouch, M., Elame, F., Abahou, H., & Choukr-Allah, R. (2017). Socio-economics and governance
of water resources in the Souss-Massa river basin. In : The Souss-Massa River Basin, Morocco,
Hdb Env Chem. Springer International Publishing, https://doi.org/10.1007/698_2016_75
Harris, J. M., Roach, B., & Codur, A. (2017). The economics of global climate change. Global
Development and Environment Institute Tufts University.
HCP. (2014). Haut-commissariat au plan. Monographie régionale 2014.
HCP. (2017). Haut-commissariat au Plan, Annuaire statistique régional Souss Massa.
HCP. (2019). Haut-commissariat au Plan, le Maroc en chiffres.
IPCC. (2019) [Special report]. Land is a critical resource, 08(August) [Press release].
Lionboui, H., Boudhar, A., Lebrini, Y., Htitiou, A., Elame, F., Hadria, R., & Benabdelouahab,
T. (2022). Digitalization and agricultural development: Evidence from Morocco. In M. Behnassi,
M. B. Baig, M. T. Sraïri, A. A. Alsheikh, & A. W. A. Abu Risheh (Eds.), Food security and
climate-smart food systems. Springer. https://doi.org/10.1007/978-­3-­030-­92738-­7_16
MADRPM. (2017). Ministère de l’agriculture, du développement rural et de la pèche maritime
Agriculture en chiffre, édition, 2018.
MADRPMDREF, agriculture en chiffres 2018, Edition. (2019).
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In Climate change:
The social and scientific construct by Suhaib A. Bandh. Springer Nature Publishing. https://
link.springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Mark, J., & Goldberg, M. A. (2001). Multiple regression analysis and mass assessment: A review
of the issues. The Appraisal Journal, 56, 89–109.
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Nisbet, R., Miner, G., & Yale, K. (2018). Handbook of statistical analysis and data mining appli-
cations (2nd ed., pp.  773–781). ISBN 9780124166325). https://doi.org/10.1016/B978-­0-­
12-­416632-­5.00022-­0
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/
book/9781774637210
Smith, G. (2018). Step away from stepwise. Journal of Big Data, 5(1), 32. https://doi.org/10.1186/
s40537-­018-­0143-­6
Tyner, W.  E., & Ouraich, I. (2014). Climate change impacts on Moroccan agriculture and the
whole economy [WIDER working paper]. 2014.46 p, April.
World Bank. (2010). Rapport sur le développement 2010 dans le monde. Développement et
changement climatique.
Chapter 15
Impact of Climate Change
on Environmental Fate and Ecological
Effects of Pesticides

Muhammad Adil, Ghazanfar Abbas, Rabia Naeem Khan, and Faheem Abbas

Abstract  Climate change refers to the long-term variations in weather conditions


observed over a period of several decades, regardless of natural or human intrusions
into a balanced earth’s ecosystem. Pesticides are extensively used for improving the
quality and yield of commodities through the prevention, mitigation or eradication
of pests, since decades. Climate change markedly affects the environmental fate of
pesticides by modifying their partitioning process between the ecological compart-
ments. It may reduce the pesticide residues on crops followed by increased suscep-
tibility to pests and diseases, thus indicating the need for more frequent pesticide
application during the forthcoming growing season. Likewise, the abundance of
pests and pathogens associated with climate change will also necessitate the plenti-
ful and repeated use of pesticides. Even the reintroduction and widespread use of
restricted or banned pesticides have also been forecasted for some countries.
Additionally, the existing pest management approaches might be ineffective for
counteracting the rapidly growing pesticide resistance under warm climatic condi-
tions. Pesticides are considered as one of the major environmental pollutants with
proven ectotoxic effects that can be further potentiated by concurrent exposure to
climate change. Therefore, the interaction of climate change and pesticide residues
necessitates thorough explication to devise preventive and legislative measures for
minimizing their combined ecotoxic impact. This chapter describes the environ-
mental fate and ecotoxic effects of pesticides in the milieu of climate change.

Keywords  Climate change · Pesticides · Environmental fate · Ecotoxicity

M. Adil (*)
Pharmacology & Toxicology Section, University of Veterinary & Animal Sciences, Lahore,
Jhang Campus, Jhang, Pakistan
e-mail: muhammad.adil@uvas.edu.pk
G. Abbas · R. N. Khan · F. Abbas
Microbiology Section, University of Veterinary & Animal Sciences, Lahore, Jhang Campus,
Jhang, Pakistan

© The Author(s), under exclusive license to Springer Nature 247


Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6_15
248 M. Adil et al.

1 Climate Change: Causes and Potential Consequences

Climate change represents the variation of climate over a long period of time (from
decades to centuries), primarily triggered by anthropogenic activities in conjunction
with natural factors (Dubey & Sudhakar, 2021). Huge emissions of greenhouse
gases resulting from the combustion of fossil fuels and deforestation are the main
drivers of progressively increasing global warming. Meanwhile, severe and consid-
erable changes in atmospheric carbon dioxide level, environmental temperature,
precipitation pattern, melting of glaciers, deterioration of polar ice caps and rising
sea levels are the typical indicators of climate change, affecting virtually all coun-
tries of the world (Delcour et  al., 2015; Gatto et  al., 2016). Therefore, climate
change is regarded as a global challenge of the current century. Certain regions of
the world including the northern Europe, central and northern Asia, and south and
north America are expected to experience increased precipitation, whereas others,
such as the southern Asia, Mediterranean countries, and southern Africa, are pro-
jected to suffer from considerable droughts (Gatto et al., 2016). Enhanced precipita-
tion and environmental temperature are forcasted to shift the production of corn and
wheat towards the northern areas of United States, whereas southern regions may
potentially suffer from reduced crop yield in view of extremely hot and dry climatic
conditions (Tubiello et al., 2002). Likewise, the recently temperate southern region
of England may become conducive for growing maize, sunflower, peaches and
grapes as a result of gradually rising temperature (Bloomfield et al., 2006; Fuller
et al., 2001).
Besides impairing the salinity, arability, nitrogen deposition and productivity of
soil, climate change has also been implicated in affecting the ecological diversity
and dynamism (Carere et  al., 2011; Gutierrez et  al., 2008; Harvell et  al., 2002;
Miraglia et  al., 2009). Climate change may alter the abundance of pests, weeds,
fungi, and other crop intruders by diminishing their overwintering mortality and
generation times (Olfert & Weiss, 2006; Malla et al., 2022; Bandh, 2022a, b; Bandh
et al., 2021, 2022, 2023; Mushtaq et al., 2020; Parray et al., 2022). Variations in
rainfall pattern, atmospheric temperature and CO2 are considered as the main trig-
gering factors of pest distribution (Gutierrez et al., 2006; Rafoss & Saethre, 2003).
Accordingly, significant variations in climatic conditions can disrupt the interspe-
cies relationship such as predation, competition, migration, parasitism and disease
transmission with consequent contamination and scarcity of food supplies (Hall
et al., 2002; Lepetz et al., 2009; Roos et al., 2011). Climatic factors, particularly,
temperature, rainfall and humidity have been implicated in determining the inci-
dence of several vector-borne disease such as cholera and malaria (Lipp et al., 2002;
Rogers & Randolph, 2000). The possible ways by which organisms respond to cli-
mate change include the shifting of climate niche along the phenological, spatial or
physiological axes (Prakash, 2017).
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 249

2 Pesticides: Historical Perspective, Classification


and Significance

2.1 Historical Aspects of Pesticides

Pesticides are chemical substances used to control or eradicate pests including


harmful arthropods (fleas, flies, lice, mites, mosquitoes, termites, and ticks), inver-
tebrates (nematodes, snails, and slugs), microorganisms (algae, bacteria, fungi, and
viruses), weeds, wild birds and rodents. Accordingly, acaricides, algaecides, insec-
ticides, avicides, miticides, herbicides, bactericides, nematicides, fungicides, mol-
luscicides, virucides and rodenticides are characteristic subtypes of pesticides
(Fig. 15.1). Although primarily meant for use on agriculture land, the application of
pesticides is also extended to other areas including gardens, railway lines and public
places (Grube et  al., 2011). As a result of increasing world population and high
demand of food supplies, the consumption of pesticides for crop protection is expo-
nentially rising.

Acaricides
Virucides Algicides

Rodenticides Avicides

Nematicides Pesticides Bactericides

Molluscicides Fungicides

Miticides Herbicides
Insecticides

Fig. 15.1  Various types of pesticides


250 M. Adil et al.

Pesticides are being utilized by humans to combat pest invasion since many
decades (Fishel, 2010). Prior to the development of synthetic pesticides, natural
compounds were used by the ancient people to control pests. Sumerians pioneered
the use of sulfur compounds as insecticides about 4500 years ago (Unsworth, 2010).
Ancient Romans used to control pests and weeds through the burning of sulfur and
application of salt, respectively (Delaplane, 2000). In contrast, mercury and arsenic
compounds from animal, plant, or mineral sources were utilized by Chinese to con-
trol lice around 3200 years ago. Usually, smoking was used for the application of
volatile substances to control insects. The hedge clippings, chaff, fish, crabs, animal
dung, or other products were burnt in such a way that a malodorous smoke was
spread to the whole vineyard, crop or orchard (Unsworth, 2010).
During the late 1800s, copper and sulfur were consumed by Swedish farmers as
antifungals for preserving fruits and potatoes (Sheail, 1991). Since then many inor-
ganic chemicals such as copper sulfate and lime arsenic are used as antifungal
agents (Bernardes et  al., 2015). The first insecticide chemical, Paris green, was
introduced in the US market in 1867 (Kogan, 1998). Later on, dichlorodiphenyltri-
chloroethane (DTT) was widely used across the world on account of its ability to
control the insect-borne diseases like typhus and yellow fever, and relatively fewer
toxic effects on mammals (Ross, 2005; Zhang et al., 2011). Nevertheless, DDT was
banned in USA after the second world war, due to its long-term bioaccumulation in
mammalian tissues, and undesirable effects on nontarget animals and plants
(Barnhoorn et  al., 2009). Accordingly, more effective and less inexpensive pesti-
cides including alderin, endrin, dieldrin, β-hexachlorocyclohexane, and
2,4-­dichlorophenoxyacetic acid were used to substitute the DDT. Nicotine, helle-
bore and pyrethrins were the early plant-derived pesticides used against aphids,
body lice and other types of insects (Fishel, 2010). In the late nineteenth century,
nicotine sulfate and calcium arsenate were consumed by Americans to control pests.
From 1970 to 1990, new families of pesticides including azolones, chloronicotinyl,
spinosyn, triazolopyrimidine, organophosphates, fiprole diacylhydrazine, strobi-
lurin, carbamates, triketone and isoxazoles were introduced in the markets
(Bernardes et al., 2015; Zhang et al., 2011).

2.2 Classification of Pesticides

Although pesticides can be categorized on the basis of mechanism of action and


target pest species, the most common classification scheme is based upon their
source and chemical nature (Buchel, 1983), and is therefore described in the current
chapter. As per this classification pattern, pesticides are categorized into synthetic
pesticides, biopesticides and nanopesticides (Fig. 15.2).
A wide range of synthetic chemicals have been used as the mainstay of pesticides
for the control and eradication of pests since decades. Synthetic pesticides are fur-
ther expanded and diversified as a result of latest innovations in the fields of drug
discovery and medicinal chemistry. Commonly used synthetic pesticides include
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 251

Fig. 15.2  Classification of pesticides on the basis of chemical nature

organophosphates, organochlorines, carbamates, pyrethroids, amides, azotic hetero-


cyclic compounds, and anilines. Organophosphates constitute a class of highly toxic
pesticides that covered almost 48.6% of all the pesticides used in 1997 (Zhang &
Zhang, 2007). Following their use for warfare purpose during the second world war,
the organophosphates have been substantially consumed in medicine, cosmetics and
agriculture (Zahran et  al., 2005; Zulin et  al., 2002). These pesticides exert their
pharmacological action by irreversibly inhibiting an enzyme called acetylcholines-
terase that carries out the degradation of acetylcholine in the nervous system of
several invertebrate and vertebrate species. Even though organophosphates can be
easily degraded, their residues can cause irreversible toxicities. Therefore, these
pesticides pose a major threat to the ecosystem and food industry (Eddleston &
Phillips, 2004). Most human cases of organophosphate intoxication occur as a result
of occupational exposure or self-poisoning (Gunnell et al., 2007). Globally, the con-
sumption of organophosphates has been restricted or banned as around 3 million
people annually suffer from acute intoxication with a mortality range of 250–370,000
(Gunnell et al., 2007; Marrs, 1993). Organochlorines are chemical pesticides con-
taining five or more chlorine items. These insecticides target the nervous system of
insects resulting in paralysis, convulsions and subsequent death. But in addition,
they have serious endocrine-disrupting effects in birds, fish and mammals.
Organochlorines such as aldrin, dieldrin and DDT are characterized by prolonged
environmental persistence for about 15 years (Ansari et al., 2021). Consequently,
most of them are banned for agricultural applications throughout the world (Larson
et al., 2019).
Carbamic acid derivatives, known as carbamates, represent a different class of
pesticides that reversibly inhibit the acetyl-cholinesterase enzyme leading to paraly-
sis and death of target parasites (Kearney et  al., 2003; Morais et  al., 2011).
252 M. Adil et al.

Carbamates usually exhibit a lower environmental persistence of about 2  weeks.


Pyrethroids are semisynthetic analogues of plant-derived pyrethrins and their insec-
ticidal activity was initially documented in 1980s. Rapid insecticidal action, high
degree of biodegradability, and minimal mammalian toxicity are the key features
accounting for the vast consumption of pyrethroids during the last 30 years (Özkara
et al., 2015). Synthetic pyrethroids including cypermethrin and permethrin can be
used on a wide range of crops and insects. Nevertheless, aquatic arthropods, fish and
mollusks are highly susceptible to pyrethroid-induced toxicity (Koureas et al., 2012).
Contrary to organochlorines, carbamates and organophosphates, the newly
developed pesticides such as amides, azotic heterocyclic compounds and anilines
are relatively less harmful (Zheng et al., 2016). Currently, amides including metola-
chlor and butachlor are frequently used, however, these pesticides have been associ-
ated with prolonged environmental persistence and mutagenic effects (Özkara et al.,
2015). Imidazoles and triazoles are nitrogen-containing heterocyclic compounds
that have given rise to almost 70% of the newly developed pesticides (Zheng et al.,
2016). Pendimethalin and trifluralin, representing the aniline group of pesticides,
are no longer used in several European countries on account of hepatic and thyroid
impairments in aquatic organisms (Özkara et al., 2015).
Increasing pest resistance and growing environmental contamination associated
with synthetic pesticides necessitated the discovery and development of more effec-
tive and eco-friendly biopesticides. Green pesticides, natural pesticides or bio-based
pesticides comprise of different ingredients such as herbal extracts, toxins, hor-
mones and pheromones (Ansari et  al., 2021). These are obtained from natural
sources and can cover a wide range of substances including microbial metabolites
and components from genetically transformed, pest-resistant crops. Certain bio-
chemical pesticides kill the insects by nontoxic substances like insect sex phero-
mones and scented plant extracts. For example, eugenol and geraniol trap the
Japanese beetle (Popillia Japonica) through their action as lures (Vargas et  al.,
2000). Besides, transgenic crops containing Bacillus thuringiensis (BT) genes are
used to produce plant-incorporated protectants such as cry proteins or lambda-­
endotoxins for immobilizing and killing the insects (Ansari et al., 2021).
Nanopesticides represent an emerging class of pesticides characterized by
enhanced efficacy, controlled delivery and minimal ecotoxic effects. Owing to the
minute size and greater surface area, the nanopesticides possess unique beneficial
characteristics in comparison with their bulk counterparts. The aqueous solubility
and dispersion of hydrophobic pesticides can be enhanced by means of nanoformu-
lation technology. Moreover, nanoencapsulation helps to maintain the required
effective concentration of pesticide by regulating its discharge rate and inhibiting
premature degradation (Ansari et  al., 2021). Nanotechnology-based systems are
used for target-specific delivery thus circumventing the overuse, wastage and resul-
tant ecotoxic impact of pesticides (Chen et al., 2015).
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 253

2.3 Significance of Pesticide Consumption

The manufacturing and utilization of pesticides is increasing every year to enhance


the yield of agriculture products and cope with continuously growing population.
The global pesticide consumption of 2 million tons per annum corresponds to insec-
ticides, herbicides, fungicides and other types as 29.5, 47.5, 17.5 and 5.5%, respec-
tively (De et al., 2014). Moreover, the global market share of pesticides (insecticides,
fungicides, and herbicides) was estimated as $84.5 in 2019 (Ansari et al., 2021). As
per the statistics of the Food and Agriculture Organization (FAO) for the year 2020,
the largest production, trade and usage of pesticides was recorded in USA, followed
by Brazil, China, Argentina and Canada, respectively (FAOSTAT, 2022). Table 15.1
indicates the data of leading pesticide-consuming countries during 2020.
Various types of pesticides are globally consumed to achieve different benefits in
agriculture and health sectors (Aktar et al., 2009). Pesticides can be effectively used
to minimize the transmission of several vector-borne diseases of humans and ani-
mals through the control or eradication of arthropods and rodents. Besides, pests
invading the internal or external body parts of humans and animals (such as hel-
minths and ectoparasites) and causing skin/tissue damage, poor growth rate and
impaired body condition can be eliminated by means of pesticides. The utilization
of pesticides is essential to yield almost one-third of the agriculture products.
Moreover, the lack of pesticide consumption in agriculture has been associated with
78%, 54%, and 32% declines in the production of fruits, vegetables and cereal
crops, respectively (Lamichhane, 2017). Accordingly, pesticides are necessarily
used in agriculture to kill or control insects and weeds and have resulted in exponen-
tial rise in agricultural productivity (Bernardes et al., 2015). The affordable, safe
and nutritious food increases the life expectancies of the consumers (Cooper &
Dobson, 2007). There would be reduced economic losses as a result of lower mor-
bidity and mortality with subsequent improvement in the production of livestock, if
pests are controlled in pastures. For example, the control cost of red-legged mites in
clover is USD 10/hector, and the control cost of insects for farmers in Australia is
USD 50/hector, but this can considerably increase the total wool yield (Ridsdill-­
Smith & Pavri, 2000). Pesticides are also used to protect the household items from
irreparable damage caused by termites, fungi or rodents.

3 Influence of Climate Change on the Consumption


and Environmental Fate of Pesticides

Climate change directly affects the soil condition such as organic matter, carbon
storage capacity and crack formation in previously arable land, as well as the appli-
cation of pesticides. The consequent modification leads to more faster and direct
passage of solutes and water from the soil surface to field drains, which can mediate
the loss of pesticides in target regions and contamination of groundwater (Tudi
254 M. Adil et al.

Table 15.1  Leading pesticide-consuming countries during 2020


S. no. Country Pesticide consumption (tons) in 2020
1. United States of America 407779.2
2. Brazil 377,176
3. China 273375.75
4. Argentina 241294.18
5. Russia 90534.96
6. Canada 78,893
7. France 65216.43
8. Australia 63416.48
9. India 61701.9
10. Italy 56,556
11. Turkey 53,672
12. Japan 51,970
13. Germany 48,002
14. Spain 43336.92
15. Mexico 41681.11
16. Colombia 36,711
17. Malaysia 36076.9
18. Ecuador 34081.3
19. South Africa 26,857
20. Philippines 24878.97
21. Ukraine 24,622
22. Poland 24,168
23. Paraguay 20164.28
24. Bolivia 19294.96
25. Viet Nam 19,154
26. Thailand 19006.56
27. Uruguay 16446.12
28. Republic of Korea 16,278
29. Bangladesh 15506.47
30. United Kingdom 14923.34
31. Kazakhstan 14492.01
32. Costa Rica 14113.7
33. Morocco 13,697
34. Pakistan 11871.92
35. Guatemala 11841.7
36. Myanmar 11748.89
37. Egypt 11352.49
38. Netherlands 11275.41
39. Peru 10630.59
40. Greece 10,475
Source: FAOSTAT (2022)
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 255

Fig. 15.3  Influence of


climate change on
environmental fate and
ecotoxicity of pesticides

et al., 2021). Increased temperature and altered CO2 concentration cause the dilu-
tion and diminished accumulation of pesticides by affecting the phenomena of pho-
tosynthesis, growth and development in plants (Delcour et al., 2015; Gutierrez et al.,
2008; Reilly et al., 2003). Besides, global warming also influences the nature, dis-
tribution and fate of xenobiotics in the environment (Tudi et al., 2021).
The environmental disposition of pesticides including volatilization, drift and
leaching processes and their decomposition occurring by means of photolysis, as
well as chemical and microbial reactions, are altered by changes in climatic condi-
tions (Tudi et al., 2021). Pesticides are directly exposed to environmental factors
including temperature, wind and rainfall which may alter their environmental dispo-
sition through increased volatilization, elevated degradation and wet deposition
(Keikotlhaile & Spanoghe, 2011; Noyes et al., 2009). Based upon the physicochem-
ical characteristics of the product, climatic conditions and the method of application
volatilization can cause an almost 50% decline in the total amount of pesticide
applied to a field. Increased temperature, high soil humidity and direct exposure to
sunlight facilitate the volatilization of pesticides, thus enhancing the likelihood of
environmental pollution (Fig. 15.3). Global warming may depreciate the concentra-
tion of pesticides in soil and water through increased volatilization (Bailey, 2004;
Benitez et al., 2006), whereas, frequent and severe rains, and storms promote their
wet deposition in the terrestrial and aquatic ecosystems (Bollmohr et al., 2007).
Moreover, the persistence and efficiency of pesticides also varies with the timing
and intensity of precipitation (Delcour et al., 2015). In contrast to areas character-
ized by low precipitation and reduced soil humidity, those with a high soil moisture
content are known to exhibit adequate hydrolysis of pesticides (Bailey, 2004). The
256 M. Adil et al.

chemical and microbial degradation of pesticides is accelerated by global warming.


Increased temperature of water has been linked with a higher rate of photodegrada-
tion of pesticides (Benitez et al., 2006). By influencing the geochemistry and min-
eralogy of soil, temperature modulates the downward transport of pesticides through
the soil into the groundwater, referred to as leaching (Bloomfield et  al., 2006;
Woodruff et al., 2009). The transport of pesticides to water bodies or other locations
primarily takes place through the mechanism of drift or runoff. Elevated tempera-
ture, heavy precipitation and frequent storms induce the drift of pesticides and sub-
sequent environmental contamination.

4 Ecotoxic Effects of Pesticides in the Context


of Climate Change

Despite their proven capacity of enhancing the agricultural production, pesticides


are known to exert notable adverse effects on humans, animals and the environment
(Liu et al., 2016; Lozowicka et al., 2014; Malone et al., 2004; Skretteberg et al.,
2015). Owing to the lack of strict control measures, large amounts of banned pesti-
cides are easily available in  local markets of several countries, thereby posing
threats to the public health and environmental well-being (Ogbeide et al., 2016).
Concurrent exposure to climate change and pesticides during the critical stages of
growth, maturation and spawning may disrupt the survival and fecundity of aquatic
species. Ectothermic species including amphibians, fish and reptiles are usually
more susceptible to the toxicological interaction of climate change and pesticides.
Moreover, the alpha predators (carnivores) at the apex of food chains exhibit con-
siderable biomagnification of pesticides owing to the processes of solvent depletion
and solvent switching (Braune et al., 2005).
The variables of climate change, predominantly temperature and humidity, sig-
nificantly influence the bioavailability, environmental fate and ecotoxic effects of
pesticides. Exposure to high temperature has been correlated with accelerated meta-
bolic conversion and enhanced toxicity of pesticides in several species of organ-
isms. For instance, the detrimental effects of atrazine (herbicide), endosulfan
(insecticide), and chlorothalonil (fungicide) were potentiated by increasing tem-
perature in green frog, catfish, rainbow trout and grass shrimp, respectively (Capkin
et  al., 2006; DeLorenzo et  al., 2009; Gaunt & Barker, 2000). Likewise, elevated
temperature led to increased pyrethroid toxicity in leopard frogs and water fleas
(Materna et  al., 1995; Ratushnyak et  al., 2005). However, elevated temperature
influences the toxicity of pyrethroids in a species-dependent manner (Narahashi,
2000; Noyes et al., 2009). These variable effects of temperature depend upon the
altered level of metabolizing enzymes as well as the toxicity profile of metabolic
products. Freshwater fish exposed to chlorpyrifos and endosulfan demonstrated sig-
nificant reduction in the upper limits of temperature tolerance (Patra et al., 2007).
Concomitant exposure to elevated temperature and chlorophenoxy herbicides per-
turbed the antioxidant defense system in soft shell clam (Greco et al., 2011).
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 257

Besides, elevated salinity increased the vulnerability of aquatic organisms to pes-


ticide intoxication by altering their osmoregulatory function (Fortin et  al., 2008;
Waring & Moore, 2004). The interactive effect of salinity and pesticide mixture
(containing chlorpyrifos and DDT) was recorded on the survival of copepod (Staton
et al., 2002). Salinity variation led to abnormal limb regeneration in male crabs by
reducing the effective concentration of permethrin (Stueckle et al., 2009). Moreover,
the induction of flavin-containing monooxygenases in the fish followed by enhanced
bioactivation of aldicarb has also been ascribed to high salinity (Wang et al., 2001).
Additionally, larval hard clams subjected to elevated CO2 (hypercapnia) and low
oxygen (hypoxia) were highly prone to resmethrin toxicosis (Garcia et al., 2014).

5 Conclusions and Recommendations

Growing human population, overexploitation of natural resources, extensive indus-


trialization and resultant rise in environmental pollution are paving the way towards
global warming and associated profound variations in climatic conditions. Prolonged
drought periods and altered patterns of environmental temperature, humidity and
rainfall are the typical consequences of anthropogenic actions (Trenberth, 2018).
Climate change is influencing virtually every form of life on the earth including
humans, animals, plants and microorganisms. Similarly, the abiotic components of
ecosystem such as air, water (including fresh and marine water bodies as well as
glaciers) and soil are also subjected to physicochemical, microbiological or toxico-
logical variations on account of climate change (Daba et  al., 2018; Jacob &
Winner, 2009).
The efficiency and throughput of agricultural practices are promptly evolving by
means of technological innovations to fulfill the increasing food demand for tre-
mendously rising human population. Despite the development of high-yielding and
pest-resistant, genetically modified, hybrid crop varieties and mechanized farming
systems, the use of pesticides is inevitable to combat the profusion and increasing
pest burden, predominantly in temperate regions. Nevertheless, the inappropriate
consumption and disposal of pesticides may lead to contamination of soil, water and
air followed by detrimental health and ecological outcomes. Moreover, the environ-
mental fate and ecotoxicity of pesticides are further confounded by simultaneous
exposure to adverse climatic conditions (Noyes et  al., 2009; Saha et  al., 2019).
Thus, anthropogenic practices contributing to climate change should be replaced by
sustainable, green industrial processes for the production of minimally hazardous,
eco-friendly and recyclable products. Afforestation and biodiversity conservation
are also requisite for controlling global warming and ecological imbalance (Díaz
et al., 2009; Reyer et al., 2009).
The environmental contamination and harmful effects of pesticides can be dimin-
ished by promoting their rational application and proper disposal. Moreover, estab-
lishing, monitoring and regulating the maximum residue limits of pesticides in
agricultural products can help to preclude the indiscriminate use of pesticides by the
258 M. Adil et al.

farmers (Kumar et al., 2019). Alternative pest eradication strategies such as inte-
grated pest management involving the minimal application of relatively harmless
pesticides in conjunction with cultural and biological control methods have also
been recommended (Zacharia, 2011). There is a growing concern regarding the
transition towards climate-resilient, organic agriculture which accentuates on the
use of crop rotation, biofertilizers, composting and biological pest control methods.
Besides, the large-scale development of useful and ecosafe, biopesticides can help
to minimize the toxicity problems ascribed to typically-used synthetic pesticides
(Nawaz et al., 2016). Finally, research studies targeting the development of efficient
and cost-effective techniques for the biodegradation of pesticide residues in organic
matter are also suggested.

References

Aktar, M. W., SenGupta, D., & Chowdhury, A. (2009). Impact of pesticides use in agriculture:
Their benefits and hazards. Interdisciplinary Toxicology, 2(1), 1–12. https://doi.org/10.2478/
v10102-­009-­0001-­7
Ansari, I., El-Kady, M. M., Arora, C., Sundararajan, M., Maiti, D., & Khan, A. (2021). A review
on the fatal impact of pesticide toxicity on environment and human health. Global Climate
Change, 361–391.
Bailey, S.  W. (2004). Climate change and decreasing herbicide persistence. Pest Management
Science, 60(2), 158–162. https://doi.org/10.1002/ps.785
Bandh, S.  A. (2022a). Climate change: The social and scientific construct (1st ed.). Springer.
https://doi.org/10.1007/978-­3-­030-­86290-­9
Bandh, S. A. (2022b). Sustainable Agriculture: Technological progressions and transitions (1st
ed.). Springer. https://doi.org/10.1007/978-­3-­030-­83066-­3
Bandh, S.  A., Shafi, S., Peerzada, M., Rehman, T., Bashir, S., Wani, S.  A., & Dar, R. (2021).
Multidimensional analysis of global climate change: A review. Environmental Science
and Pollution Research International, 28(20), 24872–24888. https://doi.org/10.1007/
s11356-­021-­13139-­7
Bandh, S.  A., Parray, J.  A., & Shameem, N. (2022). Climate change and microbial diversity:
Advances and challenges (1st ed.). Apple Academic Press. https://www.routledge.com/Climate-­
Change-­and-­Micr.obial-­Diversity-­Advances-­and-­Challenges/Bandh-­Parray-­Shameem/p/
book/9781774637821
Bandh, S. A., Malla, F. A., Qayoom, I., Mohi-ud-Din, H., Butt, A. K., Altaf, A., Wani, S. A., Betts,
R., Truong, T. H., Pham, N. D. K., Cao, D. N., & Ahmed, S. F. (2023). Importance of blue
carbon in mitigating climate change and plastic/microplastic pollution and promoting cir-cular
economy. Sustainability, 15(3), 2682. https://doi.org/10.3390/su15032682
Barnhoorn, I. E., Bornman, M. S., Jansen van Rensburg, C. J., & Bouwman, H. (2009). DDT resi-
dues in water, sediment, domestic and indigenous biota from a currently DDT-sprayed area.
Chemosphere, 77(9), 1236–1241. https://doi.org/10.1016/j.chemosphere.2009.08.045
Benitez, F. J., Real, F. J., Acero, J. L., & Garcia, C. (2006). Photochemical oxidation processes for
the elimination of phenyl-urea herbicides in waters. Journal of Hazardous Materials, 138(2),
278–287. https://doi.org/10.1016/j.jhazmat.2006.05.077
Bernardes, M.  F. F., Pazin, M., Pereira, L.  C., & Dorta, D.  J. (2015). Impact of pesticides on
environmental and human health. Toxicology Studies-Cells, Drugs and Environment, 195–233.
Bloomfield, J.  P., Williams, R.  J., Gooddy, D.  C., Cape, J.  N., & Guha, P. (2006). Impacts of
climate change on the fate and behaviour of pesticides in surface and groundwater—A UK
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 259

perspective. Science of the Total Environment, 369(1–3, 163), –177. https://doi.org/10.1016/j.


scitotenv.2006.05.019
Bollmohr, S., Day, J. A., & Schulz, R. (2007). Temporal variability in particle-associated pesti-
cide exposure in a temporarily open estuary, Western Cape, South Africa. Chemosphere, 68(3),
479–488. https://doi.org/10.1016/j.chemosphere.2006.12.078
Braune, B.  M., Outridge, P., Fisk, A., Muir, D., Helm, P., Hobbs, K., Hoekstra, P., Kuzyk, Z.,
Kwan, M., & Letcher, R. (2005). Persistent organic pollutants and mercury in marine biota
of the Canadian Arctic: An overview of spatial and temporal trends. Science of the Total
Environment, 351–352, 4–56.
Buchel, K. H. (1983). Chemistry of pesticides. Wiley.
Capkin, E., Altinok, I., & Karahan, S. (2006). Water quality and fish size affect toxicity of endosul-
fan, an organochlorine pesticide, to rainbow trout. Chemosphere, 64(10), 1793–1800. https://
doi.org/10.1016/j.chemosphere.2005.12.050
Carere, M., Miniero, R., & Cicero, M. R. (2011). Potential effects of climate change on the chemi-
cal quality of aquatic biota. TrAC Trends in Analytical Chemistry, 30(8), 1214–1221. https://
doi.org/10.1016/j.trac.2011.06.006
Chen, M., Jensen, S.  P., Hill, M.  R., Moore, G., He, Z., & Sumerlin, B.  S. (2015). Synthesis
of amphiphilic polysuccinimide star copolymers for responsive delivery in plants. Chemical
Communications, 51(47), 9694–9697. https://doi.org/10.1039/c5cc02726h
Cooper, J., & Dobson, H. (2007). The benefits of pesticides to mankind and the environment. Crop
Protection, 26(9), 1337–1348. https://doi.org/10.1016/j.cropro.2007.03.022
Daba, M., Bazi, Z., & Belay, A. (2018). Effects of climate change on soil and water resources: A
review. Environmental Earth Sciences, 8(7), 71–80.
De, A., Bose, R., Kumar, A., & Mozumdar, S. (2014). Targeted delivery of pesticides using biode-
gradable polymeric nanoparticles. Springer.
Delaplane, K. (2000). Pesticide usage in the United States: History, benefits, risks and trends, p. 20
(2007). University of Georgia College of Agricultural and Environmental Sciences.
Delcour, I., Spanoghe, P., & Uyttendaele, M. (2015). Literature review: Impact of climate
change on pesticide use. Food Research International, 68, 7–15. https://doi.org/10.1016/j.
foodres.2014.09.030
DeLorenzo, M.  E., Wallace, S.  C., Danese, L.  E., & Baird, T.  D. (2009). Temperature and
salinity effects on the toxicity of common pesticides to the grass shrimp, Palaemonetes
pugio. Journal of Environmental Science and Health, Part B, 44(5), 455–460. https://doi.
org/10.1080/03601230902935121
Díaz, S., Wardle, D. A., & Hector, A. (2009). Incorporating biodiversity in climate change mitiga-
tion initiatives. In Biodiversity, ecosystem functioning, and human wellbeing–an ecological
and economic perspective (pp. 149–166).
Dubey, I., & Sudhakar, P. (2021). Climate change, pesticides and biodiversity: A review.
International Journal on Biological Sciences, 12(1), 10.53390/ijbs.v12.i1.8.
Eddleston, M., & Phillips, M. R. (2004). Self poisoning with pesticides. BMJ, 328(7430), 42–44.
https://doi.org/10.1136/bmj.328.7430.42
FAOSTAT. (2022). Pesticides use. Retrieved August 13, 2022 from http://www.fao.org/faostat/
en/#data/RP.
Fishel, F. M. (2010). Pest management and pesticides: A historical perspective. EDIS, 2010(1),
10.32473/edis-pi219-2009.
Fortin, M. G., Couillard, C. M., Pellerin, J., & Lebeuf, M. (2008). Effects of salinity on sublethal
toxicity of atrazine to mummichog (Fundulus heteroclitus) larvae. Marine Environmental
Research, 65(2), 158–170. https://doi.org/10.1016/j.marenvres.2007.09.007
Fuller, M., Winter, M., Hopkins, A., Bol, R., & Forrest, V. (2001). An overview of the impact of
climate change on UK agriculture. Institute of Grassland and Environmental Research.
Garcia, R. N., Chung, K. W., Key, P. B., Burnett, L. E., Coen, L. D., & DeLorenzo, M. E. (2014).
Interactive effects of mosquito control insecticide toxicity, hypoxia, and increased carbon
260 M. Adil et al.

dioxide on larval and juvenile Eastern oysters and hard clams. Archives of Environmental
Contamination and Toxicology, 66(3), 450–462. https://doi.org/10.1007/s00244-­014-­0002-­1
Gatto, M. P., Cabella, R., & Gherardi, M. (2016). Climate change: The potential impact on occupa-
tional exposure to pesticides. Annali dell’Istituto Superiore di Sanita, 52(3), 374–385. https://
doi.org/10.4415/ANN_16_03_09
Gaunt, P., & Barker, S. A. (2000). Matrix solid phase dispersion extraction of triazines from catfish
tissues; examination of the effects of temperature and dissolved oxygen on the toxicity of atra-
zine. International Journal of Environment and Pollution, 13(1/2/3/4/5/6), 284–312. https://
doi.org/10.1504/IJEP.2000.002320
Greco, L., Pellerin, J., Capri, E., Garnerot, F., Louis, S., Fournier, M., Sacchi, A., Fusi, M.,
Lapointe, D., & Couture, P. (2011). Physiological effects of temperature and a herbicide mix-
ture on the soft-shell clam Mya arenaria (Mollusca, Bivalvia). Environmental Toxicology and
Chemistry, 30(1), 132–141. https://doi.org/10.1002/etc.359
Grube, A., Donaldson, D., Kiely, T., & Wu, L. (2011). Pesticides industry sales and usage. US EPA.
Gunnell, D., Eddleston, M., Phillips, M. R., & Konradsen, F. (2007). The global distribution of
fatal pesticide self-poisoning: Systematic review. BMC Public Health, 7(1), 357. https://doi.
org/10.1186/1471-­2458-­7-­357
Gutierrez, A. P., D’Oultremont, T., Ellis, C. K., & Ponti, L. (2006). Climatic limits of pink boll-
worm in Arizona and California: Effects of climate warming. Acta Oecologica, 30(3), 353–364.
https://doi.org/10.1016/j.actao.2006.06.003
Gutierrez, A. P., Ponti, L., d’Oultremont, T., & Ellis, C. K. (2008). Climate change effects on poi-
kilotherm tritrophic interactions. Climatic Change, 87(S1), 167–192. https://doi.org/10.1007/
s10584-­007-­9379-­4
Hall, G. V., D’Souza, R. M., & Kirk, M. D. (2002). Foodborne disease in the new millennium: Out
of the frying pan and into the fire? Medical Journal of Australia, 177(11–12), 614–618. https://
doi.org/10.5694/j.1326-­5377.2002.tb04984.x
Harvell, C. D., Mitchell, C. E., Ward, J. R., Altizer, S., Dobson, A. P., Ostfeld, R. S., & Samuel,
M.  D. (2002). Climate warming and disease risks for terrestrial and marine biota. Science,
296(5576), 2158–2162. https://doi.org/10.1126/science.1063699
Jacob, D.  J., & Winner, D.  A. (2009). Effect of climate change on air quality. Atmospheric
Environment, 43(1), 51–63. https://doi.org/10.1016/j.atmosenv.2008.09.051
Kearney, P. C., Shelton, D. R., & Koskinen, W. C. (2003). Soil chemistry of pesticides. Encyclopedia
of agrochemicals.
Keikotlhaile, B.  M., & Spanoghe, P. (2011). Pesticide residues in fruits and vegetables. In
Pesticides—Formulations, effects, fate (pp. 243–252).
Kogan, M. (1998). Integrated pest management: Historical perspectives and contemporary
developments. Annual Review of Entomology, 43, 243–270. https://doi.org/10.1146/annurev.
ento.43.1.243
Koureas, M., Tsakalof, A., Tsatsakis, A., & Hadjichristodoulou, C. (2012). Systematic review of
biomonitoring studies to determine the association between exposure to organophosphorus
and pyrethroid insecticides and human health outcomes. Toxicology Letters, 210(2), 155–168.
https://doi.org/10.1016/j.toxlet.2011.10.007
Kumar, A., Thakur, A., Sharma, V., & Koundal, S. (2019). Pesticide residues in animal feed: Status,
Safety, and Scope. Animal Feed Science and Technology, 7, 73–80.
Lamichhane, J.  R. (2017). Pesticide use and risk reduction in European farming systems with
IPM: An introduction to the special issue. Crop Protection, 97, 1–6. https://doi.org/10.1016/j.
cropro.2017.01.017
Larson, S. J., Capel, P. D., & Majewski, M. S. (2019). Pesticides in surface waters: Distribution,
trends, and governing factors. CRC Press.
Lepetz, V., Massot, M., Schmeller, D. S., & Clobert, J. (2009). Biodiversity monitoring: Some pro-
posals to adequately study species’ responses to climate change. Biodiversity and Conservation,
18(12), 3185–3203. https://doi.org/10.1007/s10531-­009-­9636-­0
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 261

Lipp, E. K., Huq, A., & Colwell, R. R. (2002). Effects of global climate on infectious disease:
The cholera model. Clinical Microbiology Reviews, 15(4), 757–770. https://doi.org/10.1128/
CMR.15.4.757-­770.2002
Liu, Y., Li, S., Ni, Z., Qu, M., Zhong, D., Ye, C., & Tang, F. (2016). Pesticides in persimmons,
jujubes and soil from China: Residue levels, risk assessment and relationship between fruits
and soils. Science of the Total Environment, 542(A), 620–628. https://doi.org/10.1016/j.
scitotenv.2015.10.148
Lozowicka, B., Kaczynski, P., Paritova, А. Е., Kuzembekova, G.  B., Abzhalieva, A.  B.,
Sarsembayeva, N. B., & Alihan, K. (2014). Pesticide residues in grain from Kazakhstan and
potential health risks associated with exposure to detected pesticides. Food and Chemical
Toxicology, 64, 238–248. https://doi.org/10.1016/j.fct.2013.11.038
Malla, F. A., Mushtaq, A., Bandh, S. A., Qayoom, I., Ho-ang, A. T., & Shahid-e-Murtaza. (2022).
Understanding climate change: Scientific opinion and public perspective. In Climate change:
The social and scientific construct by Suhaib A. Bandh. Springer. Nature Publishing. https://
link.springer.com/chapter/10.1007/978-­3-­030-­86290-­9_1
Malone, R. W., Ahuja, L. R., Ma, L., Wauchope, R. D., Ma, Q., & Rojas, K. W. (2004). Application
of the Root Zone Water Quality Model (RZWQM) to pesticide fate and transport: An overview.
Pest Management Science, 60(3), 205–221. https://doi.org/10.1002/ps.789
Marrs, T. C. (1993). Organophosphate poisoning. Pharmacology and Therapeutics, 58(1), 51–66.
https://doi.org/10.1016/0163-­7258(93)90066-­m
Materna, E. J., Rabeni, C. F., & Lapoint, T. W. (1995). Effects of the synthetic pyrethroid insec-
ticide, esfenvalerate, on larval leopard frogs (Rana spp.). Environmental Toxicology and
Chemistry, 14(4), 613–622. https://doi.org/10.1002/etc.5620140409
Miraglia, M., Marvin, H. J., Kleter, G. A., Battilani, P., Brera, C., Coni, E., Cubadda, F., Croci, L.,
De Santis, B., Dekkers, S., Filippi, L., Hutjes, R. W., Noordam, M. Y., Pisante, M., Piva, G.,
Prandini, A., Toti, L., van den Born, G. J., & Vespermann, A. (2009). Climate change and food
safety: An emerging issue with special focus on Europe. Food and Chemical Toxicology, 47(5),
1009–1021. https://doi.org/10.1016/j.fct.2009.02.005
Morais, S., Correia, M., Domingues, V., & Matos, C. (2011). Urea pesticides, pesticides strategies
for pesticides analysis. In M. Stoytcheva (Ed.), ISBN: 978-953-307-460-3Tech.
Mushtaq, B., Bandh, S. A., & Shafi, S. (2020). Environmental management: Environmental issues,
awareness and abatement (1st ed.). Springer. https://doi.org/10.1007/978-­981-­15-­3813-­1
Narahashi, T. (2000). Neuroreceptors and ion channels as the basis for drug action: Past, present,
and future. Journal of Pharmacology and Experimental Therapeutics, 294(1), 1–26.
Nawaz, M., Mabubu, J. I., & Hua, H. (2016). Current status and advancement of biopesticides:
Microbial and botanical pesticides. Journal of Entomology and Zoology Studies, 4(2), 241–246.
Noyes, P. D., McElwee, M. K., Miller, H. D., Clark, B. W., Van Tiem, L. A., Walcott, K. C., Erwin,
K. N., & Levin, E. D. (2009). The toxicology of climate change: Environmental contaminants
in a warming world. Environment International, 35(6), 971–986. https://doi.org/10.1016/j.
envint.2009.02.006
Ogbeide, O., Tongo, I., & Ezemonye, L. (2016). Assessing the distribution and human health risk
of organochlorine pesticide residues in sediments from selected rivers. Chemosphere, 144,
1319–1326. https://doi.org/10.1016/j.chemosphere.2015.09.108
Olfert, O., & Weiss, R. M. (2006). Impact of climate change on potential distributions and rela-
tive abundances of Oulema melanopus, Meligethes viridescens and Ceutorhynchus obstric-
tus in Canada. Agriculture, Ecosystems and Environment, 113(1–4), 295–301. https://doi.
org/10.1016/j.agee.2005.10.017
Özkara, A., Akyıl, D., Eren, Y., Erdoğmuş, S. F., Konuk, M., & Sağlam, E. (2015). Assessment of
cytotoxic and genotoxic potential of pyracarbolid by Allium test and micronucleus assay. Drug
and Chemical Toxicology, 38(3), 337–341. https://doi.org/10.3109/01480545.2014.966831
Parray, J.  A., Bandh, S.  A., & Shameem, N. (2022). Climate change and microbes:
Impact and vulnerability (1st ed.). Apple Academic Press. https://www.routledge.com/
262 M. Adil et al.

Climate-­C hange-­a nd-­M icrobes-­I mpacts-­a nd-­Vulnerability/Parray-­B andh-­S hameem/p/


book/9781774637210
Patra, R.  W., Chapman, J.  C., Lim, R.  P., & Gehrke, P.  C. (2007). The effects of three organic
chemicals on the upper thermal tolerances of four freshwater fishes. Environmental Toxicology
and Chemistry, 26(7), 1454–1459. https://doi.org/10.1897/06-­156r1.1
Prakash, S. (2017). Climate change and need of biodiversity conservation: A review. International
Journal of Applied Research, 3(12), 554–557.
Rafoss, T., & Saethre, M. G. (2003). Spatial and temporal distribution of bioclimatic potential for
the codling moth and the Colorado potato beetle in Norway: Model predictions versus climate
and field data from the 1990s. Agricultural and Forest Entomology, 5(1), 75–86. https://doi.
org/10.1046/j.1461-­9563.2003.00166.x
Ratushnyak, A. A., Andreeva, M. G., & Trushin, M. V. (2005). Effects of type II pyrethroids on
Daphnia magna: Dose and temperature dependences. Rivista di Biologia, 98(2), 349–357.
Reilly, J., Tubiello, F., McCarl, B., Abler, D., Darwin, R., Fuglie, K., Hollinger, S., Izaurralde,
C., Jagtap, S., & Jones, J. (2003). US agriculture and climate change: New results. Climatic
Change, 57(1), 43–67.
Reyer, C., Guericke, M., & Ibisch, P. L. (2009). Climate change mitigation via afforestation, refor-
estation and deforestation avoidance: And what about adaptation to environmental change?
New Forests, 38(1), 15–34. https://doi.org/10.1007/s11056-­008-­9129-­0
Ridsdill-Smith, T., & Pavri, C. (2000). Single spring spray protects pastures.
Rogers, D. J., & Randolph, S. E. (2000). The global spread of malaria in a future, warmer world.
Science, 289(5485), 1763–1766. https://doi.org/10.1126/science.289.5485.1763
Roos, J., Hopkins, R., Kvarnheden, A., & Dixelius, C. (2011). The impact of global warming on
plant diseases and insect vectors in Sweden. European Journal of Plant Pathology, 129(1),
9–19. https://doi.org/10.1007/s10658-­010-­9692-­z
Ross, G. (2005). Risks and benefits of DDT. Lancet, 366(9499), 1771–1772. https://doi.
org/10.1016/S0140-­6736(05)67722-­7
Saha, A., Ghosh, R. K., & Basak, B. (2019). Fate and behavior of pesticides and their effect on soil
biological properties under climate change scenario. In Sustainable Management of Soil and
Environment (pp. 259–288). Springer.
Sheail, J. (1991). The regulation of pesticides use: An historical perspective. Innovation and envi-
ronmental risks (pp. 38–46). Belhaven Press.
Skretteberg, L. G., Lyrån, B., Holen, B., Jansson, A., Fohgelberg, P., Siivinen, K., Andersen, J. H.,
& Jensen, B.  H. (2015). Pesticide residues in food of plant origin from Southeast Asia–A
Nordic project. Food Control, 51, 225–235. https://doi.org/10.1016/j.foodcont.2014.11.008
Staton, J. L., Schizas, N. V., Klosterhaus, S. L., Griffitt, R. J., Chandler, G. T., & Coull, B. C. (2002).
Effect of salinity variation and pesticide exposure on an estuarine harpacticoid copepod,
Microarthridion littorale (Poppe), in the southeastern US. Journal of Experimental Marine
Biology and Ecology, 278(2), 101–110. https://doi.org/10.1016/S0022-­0981(02)00325-­8
Stueckle, T. A., Shock, B., & Foran, C. M. (2009). Multiple stressor effects of Methoprene, per-
methrin, and salinity on limb regeneration and molting in the mud fiddler crab (Uca pugnax).
Environmental Toxicology and Chemistry, 28(11), 2348–2359. https://doi.org/10.1897/08-­651.1
Trenberth, K. E. (2018). Climate change caused by human activities is happening and it already
has major consequences. Journal of Energy and Natural Resources Law, 36(4), 463–481.
https://doi.org/10.1080/02646811.2018.1450895
Tubiello, F., Rosenzweig, C., Goldberg, R., Jagtap, S., & Jones, J. (2002). Effects of climate change
on US crop production: Simulation results using two different GCM scenarios. Part I: wheat,
potato, maize, and citrus. Climate Research, 20(3), 259–270. https://doi.org/10.3354/cr020259
Tudi, M., Daniel Ruan, H., Wang, L., Lyu, J., Sadler, R., Connell, D., Chu, C., & Phung,
D. T. (2021). Agriculture development, pesticide application and its impact on the environment.
International Journal of Environmental Research and Public Health, 18(3), 1112. https://doi.
org/10.3390/ijerph18031112
15  Impact of Climate Change on Environmental Fate and Ecological Effects… 263

Unsworth, J. (2010). History of pesticide use. IUPAC-International Union of Pure and Applied
Chemistry. Mai.
Vargas, R. I., Stark, J. D., Kido, M. H., Ketter, H. M., & Whitehand, L. C. (2000). Methyl euge-
nol and cue-lure traps for suppression of male oriental fruit flies and melon flies (Diptera:
Tephritidae) in Hawaii: Effects of lure mixtures and weathering. Journal of Economic
Entomology, 93(1), 81–87. https://doi.org/10.1603/0022-­0493-­93.1.81
Wang, J., Grisle, S., & Schlenk, D. (2001). Effects of salinity on aldicarb toxicity in juvenile rain-
bow trout (Oncorhynchus mykiss) and striped bass (Morone saxatilis× chrysops). Toxicological
Sciences, 64(2), 200–207. https://doi.org/10.1093/toxsci/64.2.200
Waring, C. P., & Moore, A. (2004). The effect of atrazine on Atlantic salmon (Salmo salar) smolts
in fresh water and after sea water transfer. Aquatic Toxicology, 66(1), 93–104. https://doi.
org/10.1016/j.aquatox.2003.09.001
Woodruff, L. G., Cannon, W. F., Eberl, D. D., Smith, D. B., Kilburn, J. E., Horton, J. D., Garrett,
R. G., & Klassen, R. A. (2009). Continental-scale patterns in soil geochemistry and mineral-
ogy: Results from two transects across the United States and Canada. Applied Geochemistry,
24(8), 1369–1381. https://doi.org/10.1016/j.apgeochem.2009.04.009
Zacharia, J. T. (2011). Ecological effects of pesticides, pesticides in the modern world-Risks and
benefits Dr M. Stoytcheva (Ed.), ISBN: 978-953-307-458-0. Tech, Available from: http://www
intechopen com/books/pesticides-in-the-modern-world-risks-and-benefits/ecological-effects-
­of pesticides.
Zahran, M., Abdel-Aziz, K., Abdel-Raof, A., & Nahas, E. (2005). The effect of subacute doses of
organophosphorus pesticide, Nuvacron, on the biochemical and cytogenetic parameters of mice
and their embryos. Research Journal of Agriculture and Biological Sciences, 1(3), 277–283.
Zhang, Y., & Zhang, Y. (2007). New progress in pesticides in the world. Pestic. Biochem., 12–15.
Zhang, K., Zhang, B. Z., Li, S. M., & Zeng, E. Y. (2011). Regional dynamics of persistent organic
pollutants (POPs) in the Pearl River Delta, China: Implications and perspectives. Environmental
Pollution, 159(10), 2301–2309. https://doi.org/10.1016/j.envpol.2011.05.011
Zheng, S., Chen, B., Qiu, X., Chen, M., Ma, Z., & Yu, X. (2016). Distribution and risk assessment
of 82 pesticides in Jiulong River and estuary in South China. Chemosphere, 144, 1177–1192.
https://doi.org/10.1016/j.chemosphere.2015.09.050
Zulin, Z., Huasheng, H., Xinhong, W., Jianqing, L., Weiqi, C., & Li, X. (2002). Determination
and load of organophosphorus and organochlorine pesticides at water from Jiulong River
Estuary, China. Marine Pollution Bulletin, 45(1–12), 397–402. https://doi.org/10.1016/
s0025-­326x(02)00094-­2
Index

A Climate change, 2, 4, 12, 21, 22, 37, 39, 40,


Adaptation, 19, 93, 96–101, 106, 114–126, 70, 72, 73, 81, 90–107, 114–126,
156, 168–178, 184, 186–189, 191, 193, 134–145, 150–162, 168–173, 175–177,
214, 221–231, 237 184–189, 191–194, 196–200, 208–217,
Adaptation strategies, 97–101, 114, 117, 120, 221–232, 234–244, 248–258
123, 187 Climate-smart agriculture (CSA), 91, 95–106,
Adoption, 39, 81, 97, 98, 169, 171, 173–177, 184–200, 214
185, 192, 195, 197, 213–215 Crop production impacts, 31, 142
Agricultural changes, 95–106 Crop residues, 48–60, 124
Agricultural management, 186
Agriculture, 2, 5, 12, 14–15, 21, 22, 30, 31,
37–39, 48, 52, 54, 59, 60, 70, 76, 81, D
90–107, 118, 124–126, 134, 135, 150, Decision-making tool, 197
159–161, 168, 170, 173, 186–188, Determinants, 116, 169, 175–177, 235
191–196, 198, 200, 208–217, 234–240, Disease management systems, 159, 160
249, 251, 253, 258
Agroforestry, 91, 95, 99, 102, 168–178, 187,
191, 192 E
Alternate wetting and drying, 15, 16 Ecotoxicity, 255, 257
Energy crops, 36, 70–83
Environmental fate, 248–258
B
Bangladesh, 103, 123, 134–145, 254
Bioenergy, 72–74, 76, 80–83, 99, 101, 104, 191 F
Biofertilizer, 6–7, 32–34, 39, 214, 258 Farmers, 5, 18, 32, 37, 39, 48, 49, 57, 60, 72,
Biofuels, 35, 48, 58, 70–77, 79–83, 101 73, 92, 97, 99–106, 115, 116, 118,
Biogas, 34–37, 39, 71, 73, 75, 78, 80, 83, 102 124–126, 135, 137, 142–145, 160, 161,
Biomass, 3, 12, 15, 18, 19, 34, 35, 37, 39, 168–178, 186, 188, 190–192, 195–199,
48–50, 55, 56, 58, 70–74, 77, 78, 209, 210, 212–216, 222, 223, 234, 243,
81–83, 93, 155, 169, 236 250, 253, 258
Farmers’ perception, 134–145, 171–173
Fertilizer application, 2–8
C Food security, 2, 4, 22, 32, 39, 52, 92, 94, 96, 97,
Carbon stocks, 81, 100 99, 100, 126, 161, 169, 186–189, 191,
Citrus, 234–244 192, 194, 195, 208–217, 234, 235, 237

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 265
Springer Nature Switzerland AG 2023
S. A. Bandh (ed.), Strategizing Agricultural Management for Climate Change
Mitigation and Adaptation, https://doi.org/10.1007/978-3-031-32789-6
266 Index

Fossils, 12, 30, 32, 35, 99 Nutrient, 2, 3, 5–7, 48–53, 55, 56, 58,
102–104, 123, 126, 135, 151, 152, 168,
187, 190, 215
G Nutrient cycling, 49, 58
Global energy demand, 72
Global warming, 30–32, 36, 39, 40, 70, 72, 73,
150–152, 156, 157, 224, 237, O
248, 255–257 Organic fertilizers, 2, 3, 17, 21, 32, 33, 35,
102, 196

H
Himalaya, 118, 169 P
Perception, 39, 114–126, 135, 138, 142–145,
169, 171, 172, 176
I Pest and disease management, 102, 160
Impact, 3, 4, 7, 15, 17, 20, 37, 48–55, 57, 59, Pesticides, 51, 93, 101, 152, 159–161, 174,
60, 70, 72, 73, 77, 81, 90, 95, 97, 98, 195, 208, 212, 213, 243, 248–258
100, 105–107, 114–116, 118, 119, Policy framework, 97, 98, 199
121–123, 125, 126, 134, 135, 142–144, Precision farming, 211
150–159, 161, 168, 169, 171–173, Production technologies, 11–22
175–177, 184, 185, 192, 194, 197, 199,
208–210, 212–214, 234–244, 252
India, 34, 49, 53, 57, 58, 75, 99, 118, 171, 177, R
195, 215, 227, 228, 254 Rice, 12–21, 30, 32–34, 39, 48–60, 78, 92, 95,
101–103, 154, 236
Risk reduction, 193
L
Livestock, 2, 3, 31, 32, 35, 38, 39, 48, 49, 51,
57, 60, 92, 94, 96, 100–103, 115, 124, S
191, 236, 253 Soil health, 48–60, 105, 106, 135
Synergies, 91, 96, 98, 99, 101, 107, 187,
189, 192
M
Maladaptation, 114–126
Methanogens, 13, 15, 18, 19, 32, 33 T
Mitigation, 8, 12, 21, 35, 80, 83, 93, 96–101, Technological innovation, 210, 211, 257
103, 106, 121, 168, 169, 177, 186–189,
191, 200, 214, 221–231
Multi-stakeholder systems, 100 V
Vulnerable ecosystem, 134–145

N
Natural resources, 12, 39, 90, 94, 99, 118, 126, Y
134, 135, 193, 208, 209, 212, 217, Yield, 2, 6, 15, 19–21, 33, 34, 37, 38, 48, 49,
235, 257 52, 59, 60, 74, 76, 82, 92, 93, 96,
Nitrogen oxide (N2O) emission, 2, 4, 7, 8, 12, 101–103, 126, 142–144, 152–154, 168,
15–18, 20, 21, 31, 196 186, 215, 234–244, 248, 253

You might also like