Impact of Agriculture On Soil Degradation I

Download as pdf or txt
Download as pdf or txt
You are on page 1of 398

The Handbook of Environmental Chemistry 120

Series Editors: Damià Barceló · Andrey G. Kostianoy

Paulo Pereira
Miriam Muñoz-Rojas
Igor Bogunovic
Wenwu Zhao Editors

Impact
of Agriculture
on Soil
Degradation I
Perspectives from Africa, Asia, America
and Oceania
The Handbook of Environmental Chemistry
Volume 120
Founding Editor: Otto Hutzinger

Series Editors: Damia Barceló • Andrey G. Kostianoy

Editorial Board Members:


Jacob de Boer, Philippe Garrigues, Ji-Dong Gu,
Kevin C. Jones, Abdelazim M. Negm, Alice Newton,
Duc Long Nghiem, Sergi Garcia-Segura, Paola Verlicchi,
Stephan Wagner, Teresa Rocha-Santos, Yolanda Picó
In over four decades, The Handbook of Environmental Chemistry has established
itself as the premier reference source, providing sound and solid knowledge about
environmental topics from a chemical perspective. Written by leading experts with
practical experience in the field, the series continues to be essential reading for
environmental scientists as well as for environmental managers and decision-
makers in industry, government, agencies and public-interest groups.
Two distinguished Series Editors, internationally renowned volume editors as
well as a prestigious Editorial Board safeguard publication of volumes according to
high scientific standards.
Presenting a wide spectrum of viewpoints and approaches in topical volumes,
the scope of the series covers topics such as
• local and global changes of natural environment and climate
• anthropogenic impact on the environment
• water, air and soil pollution
• remediation and waste characterization
• environmental contaminants
• biogeochemistry and geoecology
• chemical reactions and processes
• chemical and biological transformations as well as physical transport of
chemicals in the environment
• environmental modeling
A particular focus of the series lies on methodological advances in environmen-
tal analytical chemistry.
The Handbook of Environmental Chemistry is available both in print and online
via http://link.springer.com/bookseries/698. Articles are published online as soon
as they have been reviewed and approved for publication.
Meeting the needs of the scientific community, publication of volumes in
subseries has been discontinued to achieve a broader scope for the series as a whole.
Impact of Agriculture on Soil
Degradation I
Perspectives from Africa, Asia, America and
Oceania

Volume Editors: Paulo Pereira  Miriam Mu~


noz-Rojas 
Igor Bogunovic  Wenwu Zhao

With contributions by

M. F. Adame  A. Almagro  E. Argaman  A. R. Becker 


E. C. Brevik  C. B. Colman  A. P. Cuervo-Robayo 
F. A. Dadzie  R. de Faria Godoi  C. C. du Preez  E. Egidi 
D. J. Eldridge  F. Escusa  G. Eshel  D. S. Fernández 
M. Francos  M. Gong  M. d. T. Grumelli  M. Guevara 
R. R. Gutierrez  M. Hu  L. La Manna  G. J. Levy  P. Ma 
A. Maor  K. Z. Mganga  A. Molesworth  J. S. Motta 
M. Mu~noz-Rojas  P. T. S. Oliveira  D. I. Ospina-Salazar 
P. Pereira  S. Periasamy  M. E. Puchulu  M. Quintero-Angel 
M. A. Rosas  C. M. Rostagno  N. S. Santini  H. F. Schiavo 
R. S. Shanmugam  B. K. Singh  J. S. Sone  J. Stewart 
C. W. van Huyssteen  E. Volk  Y. Yu  Q. Zuo
Editors
Paulo Pereira Miriam Mu~ noz-Rojas
Environmental Management Laboratory Department of Plant Biology and Ecology
Mykolas Romeris Univerisity University of Seville
Vilnius, Lithuania Seville, Spain

Igor Bogunovic Wenwu Zhao


Faculty of Agriculture Faculty of Geographical Science
University of Zagreb Beijing Normal University
Zagreb, Croatia Beijing, China

ISSN 1867-979X ISSN 1616-864X (electronic)


The Handbook of Environmental Chemistry
ISBN 978-3-031-32167-2 ISBN 978-3-031-32168-9 (eBook)
https://doi.org/10.1007/978-3-031-32168-9

© 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
Series Editors
Prof. Dr. Damia Barceló Prof. Dr. Andrey G. Kostianoy
Department of Environmental Chemistry Shirshov Institute of Oceanology
IDAEA-CSIC Russian Academy of Sciences
Barcelona, Spain Moscow, Russia
and and
Catalan Institute for Water Research (ICRA) S.Yu. Witte Moscow University
Scientific and Technological Park of the Moscow, Russia
University of Girona [email protected]
Girona, Spain
[email protected]

Editorial Board Members


Prof. Dr. Jacob de Boer
VU University Amsterdam, Amsterdam, The Netherlands

Prof. Dr. Philippe Garrigues


Université de Bordeaux, Talence Cedex, France

Prof. Dr. Ji-Dong Gu


Guangdong Technion-Israel Institute of Technology, Shantou, Guangdong, China

Prof. Dr. Kevin C. Jones


Lancaster University, Lancaster, UK

Prof. Dr. Abdelazim M. Negm


Zagazig University, Zagazig, Egypt

Prof. Dr. Alice Newton


University of Algarve, Faro, Portugal

Prof. Dr. Duc Long Nghiem


University of Technology Sydney, Broadway, NSW, Australia

Prof. Dr. Sergi Garcia-Segura


Arizona State University, Tempe, AZ, USA

Prof. Dr. Paola Verlicchi


University of Ferrara, Ferrara, Italy

Prof. Dr. Stephan Wagner


Fresenius University of Applied Sciences, Idstein, Germany
Prof. Dr. Teresa Rocha-Santos
University of Aveiro, Aveiro, Portugal

Prof. Dr. Yolanda Picó


Desertification Research Centre - CIDE, Moncada, Spain
Series Preface

With remarkable vision, Prof. Otto Hutzinger initiated The Handbook of Environ-
mental Chemistry in 1980 and became the founding Editor-in-Chief. At that time,
environmental chemistry was an emerging field, aiming at a complete description
of the Earth’s environment, encompassing the physical, chemical, biological, and
geological transformations of chemical substances occurring on a local as well as a
global scale. Environmental chemistry was intended to provide an account of the
impact of man’s activities on the natural environment by describing observed
changes.
While a considerable amount of knowledge has been accumulated over the last
four decades, as reflected in the more than 150 volumes of The Handbook of
Environmental Chemistry, there are still many scientific and policy challenges
ahead due to the complexity and interdisciplinary nature of the field. The series
will therefore continue to provide compilations of current knowledge. Contribu-
tions are written by leading experts with practical experience in their fields. The
Handbook of Environmental Chemistry grows with the increases in our scientific
understanding, and provides a valuable source not only for scientists but also for
environmental managers and decision-makers. Today, the series covers a broad
range of environmental topics from a chemical perspective, including methodolog-
ical advances in environmental analytical chemistry.
In recent years, there has been a growing tendency to include subject matter of
societal relevance in the broad view of environmental chemistry. Topics include
life cycle analysis, environmental management, sustainable development, and
socio-economic, legal and even political problems, among others. While these
topics are of great importance for the development and acceptance of The Hand-
book of Environmental Chemistry, the publisher and Editors-in-Chief have decided
to keep the handbook essentially a source of information on “hard sciences” with a
particular emphasis on chemistry, but also covering biology, geology, hydrology
and engineering as applied to environmental sciences.
The volumes of the series are written at an advanced level, addressing the needs
of both researchers and graduate students, as well as of people outside the field of

vii
viii Series Preface

“pure” chemistry, including those in industry, business, government, research


establishments, and public interest groups. It would be very satisfying to see
these volumes used as a basis for graduate courses in environmental chemistry.
With its high standards of scientific quality and clarity, The Handbook of Environ-
mental Chemistry provides a solid basis from which scientists can share their
knowledge on the different aspects of environmental problems, presenting a wide
spectrum of viewpoints and approaches.
The Handbook of Environmental Chemistry is available both in print and online
via https://link.springer.com/bookseries/698. Articles are published online as soon
as they have been approved for publication. Authors, Volume Editors and
Editors-in-Chief are rewarded by the broad acceptance of The Handbook of Envi-
ronmental Chemistry by the scientific community, from whom suggestions for new
topics to the Editors-in-Chief are always very welcome.

Damia Barceló
Andrey G. Kostianoy
Series Editors
Preface

Agricultural soil degradation is a pervasive phenomenon related to agricultural


intensification and increasing food demand. There are some areas of the world more
vulnerable than others. For instance, semi-arid and arid areas are extremely vul-
nerable to soil degradation. Therefore, agriculture practices in these ecosystems
need to be planned carefully. Agricultural soil degradation is a consequence of
multiple pressures, such as land use changes (e.g., conversion of grasslands,
scrublands, or forests into agriculture areas), use of deep tillage methods, tractor
trafficking, and the use and abuse of agrochemicals and plastics. These practices
dramatically impact soil compaction, erosion, pollution, salinity, or acidification
[1]. Climate change is also expected to exacerbate soil degradation [2]. When soil is
degraded, several functions, ecosystem services, and habitat support are hampered.1
They lost their capacity to regulate soil erosion, floods, and climate, purify water and
supply food, fodder, and medicinal plants [3]. Food security is becoming one of the
most important issues in these turbulent times when inflation skyrockets and global
food trade are severely affected [4]. Therefore, looking at the soil as a vital resource
to our survival is essential. 95% of the food that we eat comes from soil.2 Although
this is recognized, the current picture could be more encouraging. Approximately
40% of soils in the world are degraded, which represents a loss of US$44 trillion.3 If
nothing is done, the situation can be even more dramatic, and by 2050, 90% of the
world’s soil will be degraded.4 Soil compaction decreases crop yields by 60%.5 Also,
soil erosion is accelerated about 1000 times through anthropogenic activities,
including intensive agriculture. Soil salinization makes unproductive 1.5 million

1
https://www.fao.org/global-soil-partnership/resources/highlights/detail/en/c/1539317/.
2
https://www.fao.org/about/meetings/soil-erosion-symposium/key-messages/en/.
3
https://www.unccd.int/news-stories/press-releases/chronic-land-degradation-un-offers-stark-
warnings-and-practical.
4
https://news.un.org/en/story/2022/07/1123462.
5
https://www.fao.org/3/i6473e/i6473e.pdf.

ix
x Preface

hectares of farmlands per year and has an annual cost of US$31 million.6 Soil
acidification also has substantial costs. In Australia, it is estimated that the annual
loss due to soil acidification is AUS $8.7 million per year [5]. Examples across the
world are multiple international agencies, such as the United Nations, which claim
that stopping soil degradation (e.g., erosion) is key to “save our future”.7 Soil
degradation is a severe problem in the global South, and there are several reports8
[6] that highlight that the populations are vulnerable to soil conditions and poor
harvest yield. For instance, in Africa, “More Than Half of Africa’s Arable Land ‘Too
Damaged’ for Food Production.”9 This is too dramatic to be true. It is time to act
urgently. The time to think or to wait has already passed. Sustain global food
security can no longer be at the expense of agriculture intensification. It is essential
to understand this. To understand it is key to know the causes of in different countries
for agricultural soil degradation. This was at the core of our motivation to develop
the compendium of different case studies worldwide focused on global agricultural
soil degradation in “The Handbook of Environmental Chemistry.” This project was
divided into two volumes. One focused on Africa, America, Asia, and Oceania
(Volume I) and Europe (Volume II). In the first volume, Argentina, Australia,
Bolivia, Brazil, Chile, China, Colombia, India, Israel, Kenya, Mexico, Peru, South
Africa, and the USA participated in this work.

Vilnius, Lithuania Paulo Pereira


Seville, Spain Miriam Mu~ noz-Rojas
Zagreb, Croatia Igor Bogunovic
Beijing, China Wenwu Zhao

References

1. Pereira P (2019) Soil degradation, restoration and management in a changing


environment. Elsevier, Amsterdam, p 249
2. Lal R (2020) Regenerative agriculture for food and climate. J Soil Water
Conserv 75:123A–124A
3. Pereira P, Bogunovic I, Munoz-Rojas M, Brevik E (2018) Soil ecosystem
services, sustainability, valuation and management. Curr Opin Environ Sci
Health 5:7–13
4. Pereira P, Zhao W, Symochko L, Inacio M, Bogunovic I, Barcelo D (2022)
Russian-Ukrainian armed conflict impact will push back the sustainable devel-
opment goals. Geogr Sustain 3:277–287

6
https://www.fao.org/global-soil-partnership/areas-of-work/soil-salinity/en/.
7
https://news.un.org/en/story/2019/12/1052831.
8
https://news.un.org/en/story/2021/09/1101632.
9
https://reliefweb.int/report/world/more-half-africa-s-arable-land-too-damaged-food-production.
Preface xi

5. Government of South Australia (2018) Soil acidity status report 2018 – SAMDB
NRM Region. https://cdn.environment.sa.gov.au/landscape/docs/mr/regional-
soil-acidity-baseline-2018-rep.pdf
6. DeValue K, Takahashi N, Woolnough T, Merle C, Fortuna S, Agostini A (2022)
Halting deforestation from agricultural value chains: the role of governments.
FAO, Rome
Acknowledgments

The editors appreciate the support of Prof. Damia Barcelo and Sofia Costa in
developing this book. Their encouragement was crucial to bring this project to
light. We would like also to acknowledge the support of the project No. 42271292,
funded by the National Natural Science Foundation of China.

xiii
Contents

Agricultural Soil Degradation in Argentina . . . . . . . . . . . . . . . . . . . . . . 1


Diego S. Fernández, Marı́a E. Puchulu, César M. Rostagno,
Ludmila La Manna, Analı́a R. Becker, Marı́a del T. Grumelli,
and Hugo F. Schiavo
Agricultural Soil Degradation in Australia . . . . . . . . . . . . . . . . . . . . . . . 49
Frederick A. Dadzie, Eleonora Egidi, Jana Stewart, David J. Eldridge,
Anika Molesworth, Brajesh K. Singh, and Miriam Mu~ noz-Rojas
Agricultural Soil Degradation in Peru and Bolivia . . . . . . . . . . . . . . . . . 69
Ronald R. Gutierrez, Frank Escusa, Miluska A. Rosas, and Mario Guevara
Agricultural Soil Degradation in Brazil . . . . . . . . . . . . . . . . . . . . . . . . . 97
Paulo Tarso S. Oliveira, Raquel de Faria Godoi, Carina Barbosa Colman,
Jaı́za Santos Motta, Jullian S. Sone, and André Almagro
Agriculture Soil Degradation in Chile . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Marcos Francos
Agricultural Soil Degradation in China . . . . . . . . . . . . . . . . . . . . . . . . . 153
Yang Yu, PanPan Ma, Qilin Zuo, Ming Gong, Miao Hu, and Paulo Pereira
Agricultural Soil Degradation in Colombia . . . . . . . . . . . . . . . . . . . . . . 177
Mauricio Quintero-Angel and Daniel I. Ospina-Salazar
Agricultural Soil Degradation in India . . . . . . . . . . . . . . . . . . . . . . . . . . 219
Shoba Periasamy and Ramakrishnan S. Shanmugam
Agricultural Soil Degradation in Israel . . . . . . . . . . . . . . . . . . . . . . . . . 259
Gil Eshel, Elazar Volk, Alon Maor, Eli Argaman, and Guy J. Levy
Agricultural Soil Degradation in Kenya . . . . . . . . . . . . . . . . . . . . . . . . . 273
Kevin Z. Mganga

xv
xvi Contents

Agricultural Soil Degradation in Mexico . . . . . . . . . . . . . . . . . . . . . . . . 301


Nadia S. Santini, Angela P. Cuervo-Robayo, and Marı́a Fernanda Adame
Agricultural Soil Degradation in South Africa . . . . . . . . . . . . . . . . . . . . 325
C. W. van Huyssteen and C. C. du Preez
Agricultural Soil Degradation in the United States of America . . . . . . . 363
Eric C. Brevik
Agricultural Land Degradation
in Argentina

Diego S. Fernández, María E. Puchulu, César M. Rostagno,


Ludmila La Manna, Analía R. Becker, María del T. Grumelli,
and Hugo F. Schiavo

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2 Types of Land Degradation in Argentina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3 Characteristics of Land Degradation in the Main Regions of the Country . . . . . . . . . . . . . . . . . . 7
3.1 Northwest Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2 Northeast Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.3 Pampas Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

D. S. Fernández (✉)
Grupo de Investigación en Geología Ambiental, Cátedra de Pedología, Facultad de Ciencias
Naturales e IML, Universidad Nacional de Tucumán, San Miguel de Tucumán, Argentina
SEGEMAR (Centro Regional Tucumán), San Miguel de Tucumán, Argentina
e-mail: [email protected]
M. E. Puchulu
Grupo de Investigación en Geología Ambiental, Cátedra de Pedología, Facultad de Ciencias
Naturales e IML, Universidad Nacional de Tucumán, San Miguel de Tucumán, Argentina
e-mail: [email protected]
C. M. Rostagno
Instituto Patagónico para el Estudio de los Ecosistemas Continentales (IPEEC) – CCT
CENPAT, CONICET, Puerto Madryn, Chubut, Argentina
e-mail: [email protected]
L. La Manna
Centro de Estudios Ambientales Integrados, Facultad de Ingeniería, Universidad Nacional de la
Patagonia San Juan Bosco (UNPSJB), Esquel, Chubut, Argentina
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET), Esquel, Chubut,
Argentina
e-mail: [email protected]
A. R. Becker
Departamento de Geología, Universidad Nacional de Río Cuarto, Córdoba, Argentina

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 1
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 1–48, DOI 10.1007/698_2022_917,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 17 December 2022
2 D. S. Fernández et al.

3.4 Cuyo Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23


3.5 Patagonia Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

Abstract Land degradation is a serious and widespread problem in Argentina.


Argentina is one of the largest agricultural goods producers of the world with
large-scale agricultural and livestock industries and generates a great pressure over
natural ecosystems. Main drivers of land-degradation processes are the advance of
agricultural frontier through fragile ecosystems (e.g., dry forests) and intensive and
simplified farming systems without an adequate rotation planning. Currently around
40% of the country’s lands are affected by degradation processes, a percentage that
increases to 60% if only the crop lands are considered. This situation generates a
decrease in land productivity and an increase in environmental costs due to loss of
ecosystem services. Although successful examples of conservation practices for
erosion control exists, the reality is that agricultural management practices must
consider other degradation processes that are acting in different regions. This chapter
reviews the recent studies regarding the types of land degradation and their drivers
that affect the different regions of Argentina. Proper management practices oriented
to soil conservation appear as a priority for local authorities and producers, espe-
cially in the context of climate change, which can exacerbate the negative effects.

Keywords Argentina, Deforestation, Land use change, Soil degradation processes,


Sustainable land management

1 Introduction

Argentina is one of the main exporters of agricultural products in the world along
with countries such as the USA, Brazil, Australia, Uruguay, among others. This
support the fact that agricultural production is the main economic activity of the
country, and until today, the principal driving force of the country development. In

Instituto Académico Pedagógico de Ciencias Básicas y Aplicadas, Universidad Nacional de


Villa María, Córdoba, Argentina
Centro de Investigación y Transferencia CONICET, UNVM, Córdoba, Argentina
e-mail: [email protected]
M. d. T. Grumelli and H. F. Schiavo
Departamento de Geología, Universidad Nacional de Río Cuarto, Córdoba, Argentina
Instituto de Ciencias de la tierra, Biodiversidad y Ambiente (ICBIA) CONICET-UNRC, Río
Cuarto, Córdoba, Argentina
e-mail: [email protected]; [email protected]
Agricultural Land Degradation in Argentina 3

2018, the agro-food sector, primary plus processed agricultural and livestock prod-
ucts, accounted for 58.8% of total exports [1]. The great influence of the agricultural
activity in the Argentina is based on a big natural comparative advantage for the
production of many agricultural products: 33.5 million hectares of arable land, deep
and fertile soils, fairly regular rainfall distribution, and direct access to the sea [2].
The agricultural activity has been changing through the history of the Argentina,
taking the period of 1860–1880 as starting point. Agriculture in Argentina domi-
nantly consists of cereal crops and livestock that was carried out in a traditional way
for a period of 100–120 years following patterns similar to those adopted in Europe.
Agriculture and livestock grazing expanded until the mid-twentieth century by
cultivating new lands, largely employing low-intensity agricultural practices
[3]. Due to the climatic characteristics of the plains of the country, with alternating
wet-dry periods and high-intensity rainfalls of short duration, the historical agricul-
tural practices elevated soil water and wind erosion [4]. This speeds up an adoption
and spread of no-tillage management during the 1970s [5]. Since then, there has been
a great increase in production driven by the adoption of new technologies and
changes in the form of organization of production, accelerating the process of
agriculturization [6]. Since the 1990s, the agricultural frontier has expanded to the
north-east, the north-west, and the west, moving into areas with drier climates and/or
less fertile soils [3, 7, 8]. The expansion of agriculture frontier outside the pampas
was possible due to an increase in rainfall during 1977–2005 period and the price
growth of agricultural commodities (principally soybean). Large dry forests areas
have been cleared in the Semi-arid Chaco Region of Argentina for cultivation
[9]. The cultivated area increased from about 15 to 40 million ha since
1988–2020, and bulk grain production shot up from about 27 to nearly 140 million
tons in the same period [10]. Soybean, maize, and wheat have become the primary
grain field crops in Argentina [11]. The soybean harvested area has increased
dramatically in the country from 980 ha in 1961 to 16 Million ha in 2019 [12]. Argen-
tina is the third soybean producer of the world behind the USA and Brazil [13]. Due
to this extraordinary increase in agricultural production associated to the land use
changes, Argentina faces serious problems of land degradation that are reviewed in
this chapter.

2 Types of Land Degradation in Argentina

Land degradation has been defined as a negative trend in land condition, caused by
direct or indirect anthropic actions that lead to a loss of actual or potential biological
productivity [14, 15]. Land use changes into agricultural land generally negatively
affect the soil and ecosystem services. For example, overgrazing leading to wind
erosion. Deforestation and tillage methods leading to water erosion. Salinization
triggered by water table-level rises as a consequence of irrigation. The land-
degradation processes in Argentina are triggered by rapid land-cover changes and
poor conservation policies [16–19].
4 D. S. Fernández et al.

Fig. 1 Main drivers and land-degradation processes in different regions of Argentina

Argentina presents great diversity in climate, physiography, vegetation, and land


use that lead to variations in land degradation type and intensity [20]. Main drivers of
land degradation in different regions of Argentine are shown in Fig. 1. In northwest
region the most widespread types of degradation are soil erosion, salinization, and
Agricultural Land Degradation in Argentina 5

Fig. 2 Spatial distribution of the number (a) and types (b) of land-degradation processes in arable
lands of Argentina. Data sources: affected areas (Prâvâlie et al. [21]); arable land extension (Volante
et al. [22])

aridity. In northeast region main degradation processes are water erosion and soil
compaction. In pampas region soil erosion, salinization, nutrient depletion, and
aridity are the main degradation processes. In Cuyo and Patagonia regions soil
erosion and salinization are the main degradation processes. Sometimes a combina-
tion of two or more degradation processes occurs in the same territory. In a recent
study Prâvâlie et al. [21] recognized up to three different processes acting in a same
area of Pampas Region (Fig. 2a). These authors analyze the land-degradation
footprint on global arable lands applying geostatistical techniques that are represen-
tative for identifying the incidence of five land-degradation processes: aridity, soil
erosion, vegetation decline (vegetation loss/damage), soil salinization, and soil
organic carbon decline. Figure 2b shows the result of the spatial analysis of land-
degradation processes in arable lands of Argentina and reveals that significant part of
Argentina soils suffers from combined effect of several degradation processes.
Although Argentina’s image is generally linked to the green prairies of the
Pampas, it actually may be defined as an arid country. An estimated 70% of its
area is in arid, semi-arid, or dry-sub-humid (drylands) environment, where only 12%
of the country’s water resources are located [23]. Desertification is a process that
affects drylands that consist in a loss of biological complexity and economic
productivity of a region, and involves vegetation degradation, water balance alter-
ation, and soil erosion [24]. In the case of Argentina’s dry lands, main land
degradation causes are overgrazing and land-use change, enhanced by long dry
periods [25]. In humid and sub-humid regions, main land degradation causes are
6 D. S. Fernández et al.

Table 1 Evolution of soil erosion affected area (Mha) in Argentina


Area affected based on mean
rate of soil loss
Total land area Total area Total area Moderate (2– Severe
affected by soil affected by affected by 10 t ha-1 (>10 t ha-1
-1
Year erosion wind erosion water erosion year ) year-1)
1956a 34.2 16.0 18.2 27.1 7.1
1990b 58.0 28.0 30.0 22.4 24.0
2015c 100.7 37.6 63.1 67.4 33.3
2017– 136.6 41 95.6 90.8 45.9
2019d,
e

Data sources: aInstituto de Suelos y Agrotecnia [28], bINTA-SAGyP [29], cCasas [30], dGaitán et al.
[27], eColazo et al. [31]

linked with agricultural practices used in agricultural intensive systems. Agricultural


intensification refers to two general processes: (1) changes in the vegetation diversity
(including crop species and varieties and other vegetation components such as trees,
trap crops, and weeds) and (2) changes in management practices and intensity of
production including soil amending, chemical use, tillage, and irrigation, among
others [26]. Due to the affected area and severity, the processes of soil erosion,
desertification, and physical and chemical degradation are the most important in
Argentina. Main driving factors of soil erosion in the country are deforestation and
the advance of agriculture frontier, overgrazing, and land-use change [27].
The area affected by erosion processes at national level has showed an increasing
trend since the 1950s decade. Table 1 shows the evolution of the total land area
affected by water and wind erosion at country level through studies made in 1956,
1990, and 2015. Moreover, the increment in total land area affected by erosion
processes accelerated during the period 1990–2015 by 73% with respect to the 35%
observed during 1956–1990 period. This trend agrees with the results of the
GloSEM (Global Soil Erosion Modelling) project based on RUSLE (Revised Uni-
versal Soil Loss Equation) model, that showed an increment of 41.6% of the surface
affected by erosion in Argentina during the period 2001–2012 as a consequence of
deforestation and land-use change [32]. Nevertheless, despite the increase of total
affected area by erosion, the soil erosion occurrence and level were mitigated by the
adoption of extended conservation practices in the Argentina [32]. The mean annual
soil loss due to water erosion was estimated in 6.2 t ha-1 year-1, while 12% of the
country lands suffer from severe soil loss rates (>10 t ha-1 year-1) concentrated
principally in dry lands of Patagonia, Cuyo, and Northwest regions [27]. In humid
and sub-humid areas of the country severe soil losses were reported in high slopes
lands like Tandil and Ventana ranges and rolling Pampas in Buenos Aires province,
ranges of Córdoba province, and deforested lands in Misiones province. According
to data obtained for the period 1950–2000, 33% of the country lands present a rate of
potential soil loss due to wind erosion greater than 150 t ha-1 year-1 [33]. Estima-
tions made for the 2020–2050 period have showed that wind erosion risk will remain
Agricultural Land Degradation in Argentina 7

stable compared to the 1950–2000 period [34, 35]. Nowadays, there are 41 Mha of
soils eroded by wind, of which 12.5 Mha are severely eroded. Most of the eroded
soils match with the area of high risk of wind erosion as Patagonia, the Northwest
Region, and Semi-Arid Pampas [31].
Soil salinization constitutes one of the most important soil degradation processes
and a major threat to agricultural productivity in arid and semi-arid areas of the
world. According to FAO [36] more than 70 Mha are affected by salinization
processes and at least 600,000 ha (one third of the total) of the irrigated soils suffer
from excessive salinization. The regions that were pointed out as the most affected
by soil salinization in the country are the semi-arid Chaco, the Salado River
depression, and the northwest of Buenos Aires province [37].
Complementary demands of irrigation for crops such as soybean, maize, and
wheat have increased, partly due to climatic change, but mainly due to the shallow
root development, derived from soil compaction [38]. In Argentina, agricultural soils
have decreased soil organic matter (SOM) content and aggregate stability and
increased bulk density in addition to pristine soils. Soils with less than 10 years of
continuous agriculture had 83%, 62%, and 106% of the pristine SOM content,
aggregate stability, and bulk density, respectively, while soils with 10–20 years of
continuous agriculture had 64%, 48%, and 116% of the pristine values, respectively
[11]. These findings indicate that SOM decreased by approximately 18% per decade
due to agricultural use.
The main land-degradation processes in the different regions of Argentina and
their current situation related to them are described below.

3 Characteristics of Land Degradation in the Main Regions


of the Country

3.1 Northwest Region

The Northwest region of Argentina covers an area of approximately 23 Mha and


includes the provinces of Jujuy, Salta, Catamarca, Tucumán, and Santiago del Estero
(Fig. 1). It is one of the most affected regions by deforestation during the advance of
agricultural frontier. Clearance for agriculture or cattle ranching was the dominant
land-cover change during the last two decades in this region [39, 40]. Deforestation
and cultivation in the region are carried out using heavy machinery, the burning of
remaining vegetation, and the plowing. These practices expose soil to abiotic factors,
resulting in an elevated soil vulnerability [9].
According to the Dry Chaco Forest Deforestation Monitoring Project [41], during
1976–2019 period deforested land in northwest region increased from 5.8% to
31.8% of the total region area (Fig. 3). This percentage reached 32.4% of the total
area according to the 2020 deforestation data (142,604 ha) from the early alert
system of deforestation of the federal government [42]. The 2000–2010 period
8 D. S. Fernández et al.

Fig. 3 Clearing forest temporal trend in Northwest Region of Argentina. (a) Historical spatial
distribution of the deforested areas in the different provinces of the region. (b) Evolution of the land
areas affected by deforestation in the region
Agricultural Land Degradation in Argentina 9

showed the highest rate of deforestation with 2.35 Mha, approximately 10.2% of
Northwest Region (Fig. 3). Such a vast land clearing process was triggered by two
factors: (1) the increase of global demand of soybean and (2) a 20–30% increase in
precipitation [43, 44].
Most important crops of the region are sugar cane, citrus, soybean, bean, wheat,
and maize. In dry lands area predominates soybean, bean, wheat, sorghum, and
pasture for cattle farms. Monoculture farming of bean and soybean contributes to
water erosion due to unprotected topsoil and the occurrence of intense rainfall during
summer season. High soil losses as a consequence of gully and rill erosion processes
are frequent in piedmont areas and intermountain valleys of the region [45, 46] while
in the oriental plains, sheet erosion predominates due to unrestricted cultivation or
overgrazing in relatively poor soils, previously covered by the Chaco forest [47].
In the province of Salta, land-use changes enhance runoff due to intensive
agricultural practices. The expansion of soybean crop area, with narrow crop rotation
with wheat and sorghum, has exacerbated land degradation by water erosion pro-
cesses [48]. Soil losses higher than 10 t ha-1 year-1 in area of 941,950 ha (15.3% of
the total area under study) were estimated under arable sloped (<15%) lands in the
Bermejo River Basin in the northeast of the provinces of Salta and Jujuy [49, 50].
Deforestation in Santiago del Estero province reached 3.5 Mha during the period
1976–2019. Of this total, approximately 41% took place during the 2000–2010
decade. This drastic land use change usually leads to important erosive processes,
soil organic carbon (SOC) decline, and loss of ecosystem services [51]. The effects
of land clearing on ecosystem functional attributes of the dry forest of Northwest
Region were analyzed by Volante et al. [44]. They found that once deforested, soils
significantly became more seasonal dependent than others under natural vegetation.
Such increase in seasonality is associated with a reduction of photosynthetic activity
during fallow period. Direct consequences of this biomass and vegetation shortage
can be expected on several ecosystem services such as erosion control and water
regulation, mainly due to greater exposure of bare soil, and lower biodiversity,
habitat quality, and green biomass availability for primary consumers during fallow.
In the semi-arid Chaco Region of Argentina, the total SOC stocks already decline
during the first year after deforestation, but then continue to decrease substantially
with increasing cropping time [9]. After 40 years of continuous cropping at the top
5 cm soil loss 61% SOC in addition to forest soil. The organic carbon stock content
up to 1 m depth also decreases with the land use change. Between 10 and 40 years
after deforestation, SOC loss at depth 0–30 cm was around 30%. Under different
land uses carbon stock content up to 1 m depth in Mollisols of Santiago del Estero
was reported by Osinaga et al. [48]. The organic C stock content expressed as
equivalent mass varied as follows: forest (119.3 mg ha-1) > pasture 87.9 mg ha-1
> continuous cropping more than 20-year (77.3 mg ha-1) > continuous cropping
less than 10-year (71.9 mg ha-1). Moreover, in the transition between the humid
Yungas and dry Chaco forests, in Jujuy Province, a decrease in SOC by 37% and
greater soil penetration resistance was noted in rangeland compared to deciduous
forest rangeland [52].
10 D. S. Fernández et al.

3.1.1 Soil Salinization in Northwest Region

Northwest Region contains extensive areas of salt-affected soils. Approximately


28% of the irrigated soils of Argentina are located in arid and semi-arid areas of
Northwest Region [53]. Salt-affected soils of these areas correspond to occurrence of
Aridisols, Entisols, and Alfisols with excessive irrigation, inefficient drainage, or
specific hydromorphic processes. The Chaco plain is a very flat and extensive plain
with a sub-humid to semi-arid climate, where conditions for salt accumulation in
topsoil and subsoil can be expected in most of its territory. The pervasive salt
retention of different natural vegetation formations in the dry sedimentary plains
of the Chaco plain ends with the onset of annual crop cultivation [54]. Salt retention
is no longer sustained under agriculture, as multiple paired site comparisons of
chloride profiles have revealed [55]. The rapid and intense land clearing occurred
in the plains of Northwest Region in the last two decades has produced changes in
unsaturated zone groundwater dynamics. Vadose salts that were not interacting with
plants or groundwater become dissolved and get moved closer to the surface by
rising groundwater level leading to soil and water salinization [54]. This
ecohydrological behavior was observed in the semi-arid plains of Salta and Santiago
del Estero [56, 57]. This process can be exacerbated with the irrigation in agricultural
lands located in these areas. In the Northwest region, 494,528 ha are under irrigation
where 11% of them are affected by secondary salinization [53]. Catamarca and
Santiago del Estero are the most affected with 21.2% and 11.3% of the total irrigated
area affected, respectively.
In the province of Tucumán, soil salinization is observed in the depressed plain of
the southeast where a shallow water table with saline composition (sodium bicar-
bonate type) were reported [58]. Locally, salinity problems were registered in
Mollisols (Ustolls) affected by temporary floods due to the active dynamic of the
rivers in the area during humid season [59].

3.1.2 Situation of the Arid Areas of Northwest Region

In the arid valleys of the west of the region land degradation is linked with
desertification processes like wind and water erosion, dune mobilization, aridity,
and salinization. The Santa María River Basin, a large basin that occupy part of the
provinces of Catamarca, Salta, and Tucumán, suffers from the expansion of desert-
ification during the period 1997–2012 [60]. According to Navone et al. [61], 15% of
the basin area (approximately 25,000 ha) suffer from severe problems of desertifi-
cation. Main drivers of desertification in the area are climate change (drier condi-
tions), water and wind erosion, and deforestation [62, 63].
The Puna region covers plateaus, high plains, and slopes of the Andes at
elevations between 3,200 and 6,000 m a.s.l. from Bolivian border to about 28°S. It
is an arid region with annual rainfalls between 100 and 300 mm concentrated during
summer. Features of this region are the characteristic endorheic basins with the
Agricultural Land Degradation in Argentina 11

development of saline playas. The predominant soils are Aridisols and Entisols that
are in a fragile environmental balance [64]. Regarding the area affected by the
degradation process, approximately 74% of the study area (120,000 km2) is affected
by ongoing land-degradation processes [65]. The magnitude of the different types of
soil degradation affecting the Puna region varies among provinces. Puna region of
Jujuy and Salta provinces showed a severe degree of land degradation. Soil erosion
by wind, chemical soil deterioration, and livestock rearing drive land degradation in
the dryer areas while water soil erosion became more important in the more humid
eastern part [61, 65].

3.1.3 Soil Contamination Evidences

Soil contamination evidences were recorded in some local areas of Northwest region
associated to industrial and mining activities [66–68]. Chemical soil degradation
refers to the accumulation of toxic chemicals and chemical processes which impact
on chemical properties that regulate life processes in the soil [69]. Soil salinization
and alkalinization due to wastes from pulp and paper industry were registered in a
rural area of Tucumán province [67]. Solid wastes were buried in agriculture lands
producing an increase in soil electrical conductivity and pH and a decrease in SOC
(Table 2). High concentration of heavy metals in topsoil were reported by Fernández
Turiel et al. [66] in lands around a lead smelter plant in Lastenia, province of
Tucumán. Soil and plant patterns of Pb, Cd, Ag, Zn, and Cu demonstrated the
effects of pollutant dispersion plumes over the area. The high Pb concentration in
soil (>5,000 mg kg-1) caused serious health problems in children from Lastenia
locality [70]. Soil contamination by abandoned mine tailings was reported in western
part of Northwestern Region where an intense mining activity took place [71]. The
tailings usually have elevated levels of heavy metals or other toxic elements. In
Catamarca province, high concentrations of Cu (20–1600 mg kg-1), Zn
(25–600 mg kg-1), As (5–4,900 mg kg-1), and Sn (40–1,900 mg kg-1) were
found in soils of surroundings of ancient mine tailings [72]. Other example of soil
contamination by heavy metals in adjacent lands to tailings was reported in the
abandoned mines of Concordia and La Poma in Salta province with high concen-
trations of As (200–613 mg kg-1) and Pb (230–5,000 mg kg-1) [68].

3.2 Northeast Region

The Northeast region of Argentina covers an area of approximately 29 Mha and


includes the provinces of Formosa, Chaco, Corrientes, and Misiones (Fig. 1). For-
mosa and Chaco share the Chaco forest, second largest forest region in South
America, with the provinces of Salta and Santiago del Estero of the Northwest
Region. Mean annual rainfall varies in this forest region from 450 mm to
1,200 mm and decreases from east to west, resulting in a division into the wet
12

Table 2 Comparison of the chemical properties between soil samples in a contaminated area of the Lules department, Tucumán province. Data source: Puchulu
et al. [65]
Electrical conductivity Carbonate Salts content Organic Calcium content Sodium content
Soil sample pH (dS m-1) (%) (mg l-1) matter (%) (meq 100 g-1) (meq 100 g-1)
Non-affected soil 6.68 0.70 0.79 0.04 2.24 0.17 0.19
Buried industrial 8.95 2.42 52.45 0.15 0.82 0.65 0.32
waste
Affected soil 7.99 2.05 3.76 0.13 1.42 2.25 0.80
D. S. Fernández et al.
Agricultural Land Degradation in Argentina 13

Chaco (900–1,200 mm) and dry Chaco (450–900 mm) [43]. Soybean cultivation and
cattle ranching are the most important proximate drivers of deforestation in the
Chaco, although their relative importance varies across the region on a base of the
annual rainfalls [73]. Soybean crops are currently limited to areas above 500 mm
rainfall [74]. Other important crops in the area are cotton, sunflower, and maize.
In the case of Corrientes and Misiones provinces, they present more humid
conditions with mean annual precipitations that vary from 1,200 mm in Corrientes
to more than 2,000 mm in Misiones. In consequence, native vegetation in Corrientes
is represented by a dry subtropical forest and shrubs called The Espinal, while in the
more humid areas of Misiones the Paranaense rainforest is located. Principal crops in
these provinces are yerba mate, tea, and rice. In Misiones it is also important the
forestry industry.

3.2.1 Land-Degradation Processes in Northeast Region

Chaco forest clearing for agricultural activities impacted in great manner in Chaco
and Formosa provinces. According to the Dry Chaco Forest Deforestation Monitor-
ing Project [41], during 1976–2019 period, deforested areas in the provinces of
Chaco and Formosa increased from 1.6% to 12.1% of their area. The evolution of
deforestation in the period 1997–2016 shows that until 2006, the deforested areas
were led by the province of Chaco. From this year, the province of Formosa began a
very marked increase in deforestation, surpassing Chaco province and leading the
Northeast region in the period 2012–2016 (Fig. 4). The general trend of deforestation
in the provinces of the region is downward with the exception of the province of
Formosa. During the 2010–2015 period, 188,660 ha of Chaco dry forest were
deforested in Formosa province mainly in Matacos and Patiño departments for
agricultural land use [41].
Land use change from forest to agriculture and cattle ranching is the most
important drive factor for land degradation in this region. Main land-degradation
process in Northeast Region is water erosion. Water erosion is particularly important

Fig. 4 Deforestation trend


in provinces of Northeast
region
14 D. S. Fernández et al.

Table 3 Areas affected by water erosion in Northeast region of Argentina


Area affected based on mean rate of soil loss
(ha)
Total land area affected by water Moderate (2–10 t ha-1 Severe (>10 t ha-1
Province erosion (ha) year-1) year-1)
Misiones 841,091 105,136 735,955
Corrientes 962,625 864,580 98,045
Chaco 1,776,726 1,666,308 110,418
Formosa 570,306 524,682 45,624
Total 4,150,748 3,160,706 990,042
Data source: Gaitán et al. [27]

in Misiones province where tropical rains and moderate to high gradient slopes
converge. In this province 24.5% of its area presents soil loss rates higher than
10 t ha-1 year-1, that is approximately 735,955 ha [27]. Gully and rill erosion are the
main types of water erosion that affect red soils, Ultisols and Oxisols, formed in a
hilly terrain with 5–15% slopes [75]. In Formosa and Chaco provinces sheet erosion
predominates due to agriculture advance or overgrazing in fine soils (Mollisols and
Alfisols). In the central parts of both provinces water erosion processes are affecting
soybean, sunflower, and cotton crops where severe soil losses were reported
[76]. Table 3 shows the areas affected by water erosion processes in the region.
Estimates show that a total of 4,150,000 ha present moderate to severe soil losses due
to water erosion (14.3% of the total area), where 990,000 ha present soil losses
higher than 10 t ha-1 year-1.
Agricultural expansion and intensification and overgrazing affect soil structure
altering the water and gaseous movement. Soils from Chaco and Formosa provinces
are silty with shallow water table which gives high vulnerability to soil sealing,
compaction, and waterlogging [77]. Approximately 770,000 ha are affected by soil
sealing, waterlogging, and floods in the central part of Chaco province affecting
principally cotton and sunflower crops [78]. Physical soil degradation processes in
Northeast region are a consequence of agriculturization that includes deforestation
(clear cutting, slash burning, and plowing), technological improvements with heavy
machinery use, and monoculture [79]. Total organic carbon, particulate organic
carbon, bulk density, and structural stability are the most affected soil properties
by agriculturization [80, 81]. In semi-arid areas of the province of Formosa, with
severe weather, and for 10 years of continuous agriculture, the losses of organic
carbon, total nitrogen, and light carbon have been significant [82]. In order to face
this situation long-term crop rotation was suggested as an alternative to monoculture
in order to increase chemical and physical quality of the soils of the region [83].
Land use change from natural woodlands to agriculture and livestock can alter the
hydrologic balance and salinity of soils that influence productivity and sustainable
land use [84]. Chaco forest and natural pastures naturally grow in moderate saline
soils. Land use management/change must be adjusted to this area; otherwise soil
salinization will tend to desertification. This is the case with livestock production in
Agricultural Land Degradation in Argentina 15

the south of the Chaco province where overgrazing in natural pastures usually tends
to forming bare soil patches with high salt content interspersed with non-saline soils.
Saline and sodic soils in this area cover an area of 950,000 ha [76]. In the east part of
the region with sub-humid to humid conditions irrigation is not widely used. Only
5% (117,000 ha) of the total irrigated lands of Argentina are located in this region
[85]. The affected area by salinity in irrigated lands is 8,350 ha, mostly located in
Corrientes province [53].

3.3 Pampas Region

The Pampas region is the most productive and highest-income area in the country.
The region has an approximate area of 60 Mha formed by the provinces of Córdoba,
Santa Fe, Entre Ríos, Buenos Aires, and La Pampa (Fig. 1). It can be divided into six
sub-regions on a base of climatic and geomorphological settings [86, 87]: Southern
pampa, flooded or depressed pampa, rolling pampa, sub-humid or sandy pampa,
semi-arid pampa, and Mesopotamian pampa (Fig. 5).

Fig. 5 Principal subdivisions of Pampas Region with indication (blue squares) of annual rainfall
spatial distribution
16 D. S. Fernández et al.

Pampas region is one of the largest temperate grasslands in the world with great
variability in environmental conditions. Annual average precipitation ranges from
1,100 mm in the east to less than 600 mm in the southwest, especially in spring-
summer and autumn period with more frequent deficits during summer. The average
annual temperature varies from 12.5°C in the south to 17.5°C in the northeast. The
natural biome is the grassland with dominant prairie vegetation, followed by the
steppe, belonging to the Pampean phytogeographic province [29]. It is an extensive
plain where loess sediments outcrop, partially reworked by fluvial action and
modified by pedogenesis [88]. The dominant soils are Mollisols, with smaller
areas covered by Alfisols, Entisols, Vertisols and in a very subordinate way
Aridisols, as well as in a specific places Inceptisols [89]. Among the Mollisols,
predominate Argiudolls (to the east) and Hapludolls, Haplustolls (to the west)
[90]. The aptitude of the lands is agricultural-livestock in similar proportions,
depending on whether it is elevated and stable landscapes or dunes or lowland
areas [89, 91, 92]. Pampas region has very favorable environmental and pedological
conditions for agriculture (cereals and oilseeds). Currently the main crops are
soybeans, maize, wheat, and sunflower, as well as barley, sorghum, and canola.

3.3.1 Physical Soil Degradation in Pampas Region

Before the conversion to agricultural land, this region was covered by grasslands
without large herbivores and a low human population. Agriculture and livestock
began to develop slowly from the arrival of immigrants with an increase in the
human population. Natural grasslands were grazed until the last quarter of the
nineteenth century when forage crops and mercantile crops began to grow. Since
then, agriculture expands the region and nowadays represents the dominant activity
[93]. The expansion of the agricultural and livestock frontier in Argentina and the
use of agricultural technology are factors that explain the increase in biological and
economic productivity of rural sector in the last five decades [94]. Historically,
farmers have made business decisions based on an economic relationship between
benefits and costs, and has generally ignored the relationship between the economic
benefit and the environmental cost of such a decision [95].
The increases in gross production in the pampas prairie were marked by an
expansion over new lands until the 70s and 80s [96], and from then on, the
productive jump can be explained by a more intensive agricultural practices. Since
the 70s, there has been an important increase in the area under arable crops, with the
cropped area increasing relative to the pasture area at an annual rate of 4% [97]. In
recent years, there has been an extreme simplification of the Pampean production
systems with a gradual replacement of traditional rotations by monoculture. This
trend toward the production of a single crop had an unfavorable impact on the soil
functions and the agroecosystem sustainability, such as the loss of organic matter
and clay by water erosion [98]. In different sectors of this region, as a response to
monoculture, physical degradation of the soil is observed [99] that strongly induces
wind and water erosion processes [35, 91, 100–105].
Agricultural Land Degradation in Argentina 17

Fig. 6 (a) Aerial photograph of soil water erosion processes in the locality of Río Cuarto, Córdoba
province. (b) Rill erosion in soils of the locality of General Cabrera, Córdoba province. (c) Gully
erosion in soils of the locality of General Deheza, Córdoba province. (d) Laminar and rill erosion in
soils of the locality of General Cabrera, Córdoba province. Photos: (a) M Cantú, November 2011.
(b) A Becker, September 2017. (c) Analía Becker, October, 2018. (d) Analía Becker, April 2019

For these reasons, the most relevant soil degradation process is wind and water
erosion. Long history of conventional tillage promotes sediment movement and loss
of the surface horizon can usually be observed [106] (Fig. 6). The widespread
adoption of no-tillage in the last two decades has slowed down the deterioration
process [107]. However, in areas under monoculture, mainly soybean, no-tillage
method was not enough to reverse soil degradation [108], although the inclusion of
cover crops is mentioned as an alternative in no-tillage systems [109–111].
According to Casas [5] in the last three decades, Buenos Aires province has
greater intensity of agricultural land use and a greater concentration of livestock
(Feedlot). The lack of rotations, the outsourcing of land use, and the excessive
demand for stubble by concentrated livestock farming contributed to the presence
of a new scenario where the processes of soil degradation are showing a different
dynamic. In general, compaction and structural deterioration processes are being
developed that help reduce infiltration, percolation, and retention of water in the soil.
Consequently, in flat areas, surface waterlogging tends to increase, and in undulating
areas surface runoff and water erosion increase [112]. In areas of the west and
southwest of the Pampas region, where the occurrence of drought periods is fre-
quent, the processes of wind erosion increase. According to Casas and Puentes
[113], the province of Buenos Aires has 35.5% of their lands affected by water
erosion. In the rolling pampa, with slopes that varies from 0.3 and 3%, highly
18 D. S. Fernández et al.

evolved Typic Argiudolls develop on loess material [91]. In this area, processes of
water erosion, degradation of the surface horizon due to loss of organic matter,
compaction and acidification due to continuous agriculture are observed.
Naturally or anthropogenic-induced compacted layers that present difficulties for
root growth, tillage, water and gaseous movement, and engineering uses were
described in the Pampean region [112]. In addition, these authors [112] point to
the occurrence of fragipan distribution in the soils on approximately 500 km2 in
some regions.
In the southwest of Buenos Aires, compaction processes are recorded that could
be attributed to the loss of macroporosity in soils under no-tillage management [114]
since this system has not been able to reverse the problems of physical degradation.
Important compaction processes were reported by Agostini et al. [115] in soils with
high organic matter (OM) content and loamy texture in the southeast of Buenos
Aires province. Soils with physical degradation processes due to compaction were
reported in the rolling pampas in Buenos Aires province due to the high silt content,
long agricultural history and their management, with a predominance of no-tillage
management [116]. In cultivated fields of the sandy Pampas, sectors with chlorotic
less developed plants are observed called “overos” or “spotted” fields. Spotted fields
occur mainly due to severe compaction at variable depths caused by cementing of
silica or Fe, very hard Bt horizons, high apparent density, or excessively alkaline
reaction [117].
The central-west part of Córdoba province is affected by processes of water
erosion, mainly linked to change and intensification of land use in fragile environ-
ments [118] where the simplification of rotations and soybean monoculture contrib-
uted to increase in the erosive process. The gullies represent one of the most severe
forms of erosion due to their impact on road infrastructure and due to their high
sediments production [119]. Migration of peanut production (3,500 ha sown in 1998
to 116,000 ha in 2012) to south-central part of province increased susceptibility of
soils and frequency occurrence of phenomenon of “blast” (dust flying by wind),
particularly in dry years [120]. Waterlogging affects a total area of 1.7 Mha in the
province of Córdoba, of which 880,000 ha correspond to agricultural land, and the
rest to natural shallows suitable for livestock use or wetlands (usually used for
biodiversity reserve and flood control).
Soils under agricultural and livestock use in south-central Córdoba province
present anthropogenic-induced evidence of compaction [121] where reduced infil-
tration favoring surface runoff and water erosion that negatively affect the behavior
of crops in the area. These soils present natural susceptibility to compaction due to
high silt and low organic matter content [121, 122]. In the current depressed areas of
south eastern Córdoba province, the presence of subsoil compacted layers, together
with the climate and the relief is the cause of great periodic floods and long period of
permanence of the surface water. It is important to mention that fragipans interfere
regionally in the dynamics of water infiltration when they are close to the
surface [123].
Water erosion processes in Mesopotamian pampa are manifested with greater
intensity in the western half, extreme south west of the Entre Ríos province, and in
Agricultural Land Degradation in Argentina 19

part of the south east sector. Water erosion affects 42% of the province with soil
losses higher than 2 t ha-1 year-1 [27]. This means that around 4.5 Mha can be
eroded and therefore must be managed using conservation practices. To mitigate soil
degradation caused by water erosion, conservation tillage systems were developed
and adopted, mostly reduced and no-tillage. However, in some regions, depending
on the type of soil, slope and intensity of rainfall, the use of conservation tillage
systems is insufficient to mitigate the erosion. Therefore, farmers and land managers
have to implement technologies such as systematization of land and construction of
terraces to control water erosion [124].
Soil compaction problems in center and north Santa Fe province were reported by
Imhoff et al. [125], whose causal agents would be silt high content and organic
matter loss that is manifested in high bulk density (>1.3 g cm-3) in topsoil. Entre
Ríos province under Vertisols used mainly for rice cultivation is highly susceptible
to compaction due to high content of expandable clay (>40%). Moreover, harvest is
often performed under saturated conditions by heavy harvesters [126].
Currently, in La Pampa province water erosion has a greater magnitude in the
central and south east zone, occupied by the natural ecosystem of the Caldenal and
dedicated to extensive cattle ranching. However, there is pressure from agriculture to
force clearings and transform extensive silvopastoral systems into agricultural ones.
In the zones affected by wind erosion like the east of La Pampa province the area
occupied by highly susceptible soils usually exceeds 50% [127]. Approximately
51% of the total area of the provinces of the region is affected by soil erosion
according to the latest estimates, with more than 8.8 Mha with soil loss rate higher
than 10 t ha-1 year-1 (Table 4). In north and center of La Pampa province studies
carried out by Duval et al. [128] indicate that agricultural use has generated a
non-critical compaction of soils at expense of 12% decrease in macropores. In
relation to soil type, Argiudolls are vulnerable to agricultural practices intensification
increasing degradation processes by compaction. Regardless of soil type,

Table 4 Soil erosion affected areas (Mha) in Pampas Region


Area affected based on
mean rate of soil loss
Total land area Total area Total area Moderate (2– Severe
affected by soil affected by affected by 10 t ha-1 (>10 t ha-
-1
Province erosion wind erosiona water erosionb year ) 1
year-1)
Buenos 17,800,285 6,987,000 10,813,285 13,918,068 3,882,217
Aires
Santa Fe 3,429,567 – 3,429,567 3,123,831 305,736
Córdoba 13,355,133 4,764,000 8,591,133 10,177,550 3,177,583
Entre 3,314,134 – 3,314,134 2,277,486 1,036,648
Ríos
La 4,034,381 2,498,162 1,536,219 3,606,315 428,066
Pampa
Total 41,933,500 14,249,162 27,684,338 33,103,250 8,830,250
Data sources: aCasas [30], bGaitán et al. [27]
20 D. S. Fernández et al.

agricultural activity generated decreases in total organic carbon (TOC) levels,


negatively affecting most of physical properties evaluated with increases in bulk
density and decrease in total porosity, mainly due to lower volume of macropores
which decreased 12% on average.
In the 1990s, the agricultural intensification advanced towards simplified produc-
tion schemes under no-tillage cropping, with spring-summer species, especially
soybean and maize. As a consequence, concern about negative nutrient balances in
the Pampas and the importance of considering these deficits in land management and
fertilization decisions have been expressed [129, 130]. Studies in croplands of the
Argentine Rolling Pampa demonstrated that the technology (reduced tillage,
no-tillage) was unable to maintain existing soil organic matter stocks with losses
of 15–20% after 30 years [108, 131].

3.3.2 Salt-Affected Soils in Pampas Region

The Pampa region contains several areas with salt-affected soils as a consequence of
waterlogging in lowlands or the presence of a shallow saline water table. According
to Taboada et al. [132], there are 11.6 Mha of soils with some degree of alkaliniza-
tion/sodification in the region. These soils are located principally in the meridional
lowlands of the north of Santa Fe province, in depressed pampa in the center of
Buenos Aires province and in the south-east of Córdoba province. Irrigation plays a
key role in the intensification of agricultural systems of sub-humid areas. In the case
of the Pampas Region, agricultural lands use pressurized irrigation systems. Soil
salinization problems were reported in irrigated crop lands as a consequence of the
use of sodic bicarbonate type groundwater for irrigation purposes [133]. The use of
sodic bicarbonate water produces an increment of soil pH and ESP (exchangeable
sodium percentage) with little or no variation of electrical conductivity. As a result,
soil sodification is the highest risk for these areas [134]. According to statistical data,
the region has more than 500,000 ha of crop lands under irrigation where 163,000 ha
(29.3%) are affected by soil salinization [53]. The province of Buenos Aires is the
most affected part of the region with 40% of the irrigated lands with soil salinization
evidence (Table 5).

Table 5 Irrigated and salt-affected soils area in the different provinces of Pampas region
Irrigated land Area affected by soil Area affected by soil
Province area (ha) salinization (ha) salinization (%)
Buenos Aires 317,720 128,560 40.4
Córdoba 102,000 17,710 17.3
Santa Fe 62,508 6,000 9.5
Entre Ríos 69,000 10,350 15.0
La Pampa 4,600 460 10.0
Total for the 555,828 163,080 29.3
region
Data source: Sánchez et al. [51]
Agricultural Land Degradation in Argentina 21

3.3.3 Agrochemical Contamination in Pampas Region

The Pampa region is characterized by no-tillage management and intensive use of


agrochemicals. Glyphosate (N-phosphono-methylglycine) is the most commonly
used herbicide in transgenic soybeans whose main metabolite, due to microbial
degradation, is aminomethyl-phosphonic-acid (AMPA). In maize crops and
soybean-maize crop-rotation sequence Atrazine (2 chloro 4 ethylamino 6 isopropyl
amino 1,3,5 triazine) is used in conjunction with Glyphosate in agricultural lands of
Argentina. Glyphosate (GLP), AMPA, and Atrazine (ATZ) residues were detected
in the upper 10 cm of soils in different provinces of Pampa region [135–138]. A
summary of several studies about GLP, AMPA, and ATZ residues in soils of the
Pampas region is shown in Table 6. In cultivated soils, GLP was detected in
concentrations between 8 and 5,200 μg kg-1, while AMPA concentration ranged
from 3 to 6,550 μg kg-1. The GLP and AMPA levels detected in soils of the
provinces of Buenos Aires and Santa Fe were not significantly different having
concentrations that ranged from 2 to 1,500 μg kg-1. The province of Córdoba has
showed remarkable differences between different areas, while the highest concen-
trations were measured in the province of Entre Ríos with an average concentration
of 2,299 μg kg-1. ATZ residues were measured in Buenos Aires and Córdoba
provinces in concentrations between 4 and 66 μg kg-1. The difference in herbicide
concentrations could be related to some physicochemical properties of the soils, like
organic matter and clay content, and different types of crops.
Some papers dealt with the study of the environmental fate of GLP and AMPA in
surface water, groundwater, and suspended particulate matter [137, 141, 143]. The
major risk of propagation is linked with runoff and the detachment of soil particles
according to studies carried out in Buenos Aires province. In suspended particulate
matter, GLP was found in 67% while AMPA was present in 20% of the samples,
while in stream sediments were also detected in 66% and 88.5% of the samples,
respectively [141]. In the surface water studied, the presence of GLP and AMPA was
detected in about 15% and 12% of the samples analyzed, respectively. Similar
results were found in the Tapalqué stream basin in the province of Buenos Aires
where GLP and AMPA residues were detected in sediments and surface water
[138]. Conversely, studies on GLP concentrations in groundwater indicate a lower
frequency of detection on shallow aquifers than in surface water [144]. Until now the
presence of GLP residues in deep aquifers is still undetectable [136].
Soil wind erosion is other potential transport process of GLP molecules in
agricultural soils. Measurable concentrations of GLP and AMPA were detected
also in the dust emitted by soils of the central semi-arid region of Argentina,
12 months after GLP application [143]. This indicates that GLP and AMPA can
potentially be a source of air contamination in windy regions.
22

Table 6 Summary of GLP, AMPA, and ATZ contents in soil samples (0–10 cm soil layer) for the provinces of the Pampas region
Buenos Aires Buenos Aires Buenos Aires Buenos Aires
Province (n = 58)a (n = 6)b (n = 5)c (n = 16)d Córdoba (n = 58)a Córdoba (n = 7)e Santa Fé (n = 58)a Entre Ríos (n = 17)f
Locations Coronel Suárez, La Olavarría, Tapalqué Quequén Grande Balcarce, Necochea, B° Ituzaingó, Piedmont area of Hersilia Urdinarrain
Plata River Basin San Cayetano, Tres Malvinas Sierras Chicas Range
Arroyos Argentinas, Marcos
Juárez, Brinkmann
Prevalent Soybean-wheat, Soybean, wheat, System barley- Soybean-wheat, sun- Soybean-maize Soybean, maize Wheat Soybean, wheat,
crop type horticultural maize, oats soybean flower-wheat- sunflower, maize
soybean
Dominant Mollisols Mollisols, Alfisols Mollisols Mollisols Mollisols Mollisols Mollisols Vertisols
soil type
Herbicide GLP AMPA ATZ GLP AMPA ATZ GLP AMPA ATZ GLP AMPA ATZ GLP AMPA ATZ GLP AMPA ATZ GLP AMPA ATZ GLP AMPA ATZ
(μg kg-1)
Mean 216 <LoD 12 28 270 – 66 231 – 401 686 – 97 297 14 1,100 1,453 – 214 <LoD – 2,299 4,204 –
Min 184 <LoD 6 2 3 – 8 7 – 35 299 – 28 80 4 130 910 – 104 <LoD <LoD 450 1,250 –
Max 248 <LoD 17 176 712 – 179 472 – 1,502 1,052 – 233 732 66 3,700 2,500 – 323 <LoD 7 5,200 6,550 –
Median 216 <LoD 12 – – – 39 201 – 232 730 – 67 223 5 285 1,200 – 214 <LoD – – – –
LoD limit of detection
a
Alonso et al. [139]
b
Pérez et al. [138]
c
Lupi et al. [140]
d
Aparicio et al. [141]
e
Bento et al. [142]
f
Primost et al. [136]
D. S. Fernández et al.
Agricultural Land Degradation in Argentina 23

3.4 Cuyo Region

The Cuyo region of Argentina covers an area of approximately 40 Mha and includes
the provinces of San Luis, Mendoza, San Juan, and La Rioja (Fig. 1). It is an
extensive area consisting from sub-humid plain in the east to semi-arid and arid
Andean regions in the west. The aridity index, defined as the ratio of potential
evaporation to precipitation [145], varies from 30 to 5 east-west [146] showing the
vulnerability to desertification processes of the entire region.
Land use change in agricultural presents different trends among the provinces of
the region. San Luis and La Rioja have the largest areas affected by deforestation. In
contrast, in Mendoza and San Juan deforested areas are not significant. According to
the Dry Chaco Forest Deforestation Monitoring Project [41], the evolution of
deforestation in the period 1997–2019 shows that the deforested areas in this region
were led by the province of San Luis (Fig. 7). The highest deforestation rate in Cuyo
region was observed during the 2012–2016 period, while deforestation trend in San
Luis province remains on rise until 2019.
Principal crops of the region are vineyards, olive, grape, maize, sorghum, and
pasture for cattle farms. Soils of the region are represented principally by Entisols
and in driest areas by Aridisols. They exhibit variations in their capacity to sustain
agricultural activities, always within the overall low productivity that characterizes
dryland soils. These variations are principally related to differences in chemical
compositions, to local combinations of climate and geomorphology and with irri-
gated agriculture [24].
The main drivers of desertification in this region are land use change and
overgrazing. Approximately 25% of the affected lands by desertification processes
in Argentina are located in Cuyo Region [147]. As a consequence of the overgrazing
in eastern undulating plains the original grassland savannah with isolated woody
plants called “El Caldenal” was replaced by unpalatable grasses [25]. Wind and
water-sheet erosion are the main soil degradation processes due to the lack of
vegetation cover. Soil degradation in the area includes the loss of biodiversity
and/or ecosystem services, water and wind erosion, and salinization [148].

Fig. 7 Deforestation trend


in provinces of Cuyo Region
24 D. S. Fernández et al.

Table 7 Soil erosion affected areas (Mha) in Cuyo region


Area affected based on
mean rate of soil loss
Total land area Total area Total area Moderate Severe
affected by soil affected by affected by (2–10 t ha- (>10 t ha-
-1
Province erosion wind erosiona water erosionb 1
year ) 1
year-1)
Mendoza 5,214,588 148,827 5,065,761 2,900,045 2,314,543
San Juan 5,120,237 942,000 4,178,237 2,255,396 2,864,841
San Luis 1,998,331 675,000 1,323,331 1,572,351 425,980
La Rioja 3,077,822 605,000 2,472,822 1,979,487 1,098,335
Total 15,410,978 2,370,827 13,040,151 8,707,279 6,703,699
Data sources: aCasas [30], bGaitán et al. [27]

Land use conversion, particularly from the second half of the twentieth century,
increases over-exploitation of natural resources, producing a notable modification of
the vegetation cover [149]. As a consequence, water and wind erosion and the
advance of the dunes affect important areas of Cuyo Region. Approximately 45%
of the total lands of the region show values of erodible fraction of 75–100%
according to results of wind erosion models applied for Argentina [135]. Approxi-
mately 38% of the total area region is affected by soil erosion according to the latest
estimates, with more than 6.7 Mha with soil loss rate higher than 10 t ha-1 year-1
(Table 7).
Livestock create pressure on soils, generating visible signs of overgrazing in most
cases. Cattle ranching is concentrated in the plains of the center-east and south of the
region and in some privileged mountain valleys with more humid conditions.
Overgrazing consequences on soil physical conditions were assessed in the El
Caldenal semi-arid area of the region. The lack of vegetation in grazed areas is
always translated into the disappearance of the topsoil layer. Reduction of topsoil
depth and rise of soil bulk density were reported in this area [150].

3.4.1 Soil Salinization in Cuyo Region

Soil salinization is a major problem in the Cuyo region. It can be triggered in


non-irrigated areas by land use changes when perennial deep-rooted native vegeta-
tion is replaced by annual crops or by excessive irrigation and deficient drainage in
irrigated lands [151]. The Cuyo region presents the largest agricultural area under
irrigation of the country. According to national statistics 500,000 ha are under
surface irrigation and groundwater irrigation [152] of which 28.3% are salt-affected
soils. The analysis of the results of each of the constituent provinces of the Cuyo
Region shows that Mendoza has 26.4% of soils under irrigation with some degree of
salinization (Table 8). The North Oasis of Mendoza province is the most affected
area by soil salinization [36]. The process is particularly intense in north and central
parts of the Mendoza river basin where soil electrical conductivity (EC) can reach
Agricultural Land Degradation in Argentina 25

Table 8 Irrigated and salt-affected soils area in the different provinces of Cuyo region
Irrigated land Area affected by soil Area affected by soil
Province area (ha) salinization (ha) salinization (%)
Mendoza 276,324 73,213 26.4
La Rioja 51,738 5,300 10.2
San Juan 95,704 53,830 56.2
San Luis 76,437 9,580 12.5
Total for the 500,203 141,923 28.3
region
Data source: Sánchez et al. [51]

16 dS m-1 [153]. However, the EC of non-cultivated soil is significantly higher in


addition to cultivated soil. Satellite image analysis of cultivated and non-cultivated
lands with the respective salinity maps revealed that 97% of the cultivated area had
EC lower than 6 dS m-1 and the whole cultivated area had EC values lower than
8 dS m-1 on the average [154]. Main causes of soil salinization in the area are the
excessive irrigation and the high level of a saline water table with EC values in range
from 3 to 3.5 dS m-1 [155]. The salinity of the soils in all the irrigated oasis of the
province of Mendoza was surveyed by Vallone et al. [156] who determined that in
the Tunuyán River basin 25% of the soils have EC values greater than 4 dS m-1.
Moreover, an increase in chloride toxicity has been noted in vineyards in the
Mendoza River Basin [157]. Possible reasons for this situation are the use of
chloride-sensitive rootstocks, an increase of the drip irrigation area, irrigation man-
agement strategies with water deficit, deterioration in the groundwater quality, and
poor soil physical condition (compaction, cementing, poor internal drainage, and
poor infiltration) [158].
San Juan province shows a dramatic situation where 56.4% of all irrigated soils
were affected by salinization processes (Table 8), due to excessive irrigation that
lead to water table rise [53].
In non-irrigated areas the replacement of natural dry forest for crops causes
phreatic level rises and the transport of salts to soil surface [159]. In some sectors
of San Luis Province (near Villa Mercedes and Juan Jorba cities), phreatic rises of up
to 10 m were observed in the last three decades and, in some cases, surface outcrops
of groundwater with high salinity [160].

3.5 Patagonia Region

Patagonia is the southernmost region of South America. The part located in Argen-
tina extends from the south of the Colorado River, from about 36°S latitude, to the
Beagle Channel in the southern part of the Tierra del Fuego Island located at 55°S
latitude. Longitudinally, it extends between the Andean range and the Atlantic coast.
It is mostly characterized by different levels of plateaus and mountains interrupted
26 D. S. Fernández et al.

by transversal valleys of the main rivers and numerous close basins. The north-south
distribution of the Andes range acts as a barrier for the humid winds coming from the
west giving rise to a narrow, humid strip characterized mainly by forests, that, in few
kilometers changes into dry lands with highly evaporative conditions, that charac-
terize the Patagonian rangelands [146].
The Patagonian rangelands represent approximately 95% of the territory of
Patagonia (about 790,000 km2) and are characterized by an arid to semi-arid climate,
with annual precipitation in most of its extension below 400 mm. Central portion of
Patagonia in most parts has less than 200 mm of rain per year. In a narrow strip of the
west Patagonian rangelands (the ecotone with the woodland biome), and in the north
east portion and many parts of the coastal area mean annual precipitation is above
250 mm per year reaching up to 600 mm year-1. The strong, constant west winds
(westerlies) are dominant across the region [161]. The strong winds as well as some
intense rainfall events are the main geomorphic factor explaining the erosion pro-
cesses in the Patagonian rangelands.

3.5.1 Agricultural Land Use Impacts on Soil Degradation: Erosion


Processes in Patagonia

Sheep raising for wool and meat production has been the main economic activity.
Since 1885, settlement and economic development were based on sheep ranching.
Almost free of domesticated animals in 1885, by 1910, the Patagonian rangelands
supported nearly 12 million sheep [162]. In 1952, it attained its maximum stocking
rate of approximately 25 million [163]. Goat farming is practiced in small areas of
the Patagonian rangelands, mainly in the northwestern region. The high environ-
mental spatial heterogeneity as well as the annual and inter-annual rainfall variability
affects the spatial and temporal uncertainty in forage production and has favored the
negative impact of livestock grazing. The overestimation of the rangeland carrying
capacity favored overgrazing and land degradation. Thus, in the rangelands of
Patagonia, continuous sheep grazing, practiced for more than a century, has changed
the vegetation structure, mainly its cover, and accelerated soil erosion [164]. Indeed,
soil erosion is recognized as the main degradation process underlying the desertifi-
cation that affects a large proportion of the arid and semi-arid lands of Patagonia
[165]. In spite of its aridity, water erosion seems to have been the most important
process of soil degradation [166, 167]. In some areas, wind erosion is an important
geomorphologic process and seems to have modeled the landscape before the
introduction of domestic livestock. However, sheep grazing has contributed to an
acceleration of the wind erosion process and has activated some originally fixed sand
dunes.
Desertification has been recognized as a serious problem affecting the arid and
semi-arid lands of Patagonia. In this region, the severe and very severe conditions
affect ~30% of these rangelands [165]. One of the main land-degradation processes
underlying the desertification syndrome in this region is soil erosion. Desert pave-
ments, a common feature of many naturally eroded areas, are widespread all over
Agricultural Land Degradation in Argentina 27

Patagonia and have been considered a good indicator of accelerated soil


erosion [168].
Soil erosion is a natural process that can be accelerated by different economic
activities like sheep grazing, oil and gas exploration, or some events such as natural
or human-induced fires. The erosion process can remove nutrient rich fine soil
particle and further deplete already poor soils [167, 169]. Thus, in the Patagonian
steppe, as in many arid and semi-arid rangelands, soil erosion can profoundly affect
the capacity of the land to produce different goods and services. Even in the
sub-humid rangelands from the west, where soil fertility and climatic conditions
are more favorable than in other areas of Patagonia, soil erosion is an important
process of soil degradation [170]. Although in the Patagonian rangelands wind is an
important geomorphological agent that in some localized areas has deeply modified
the landscapes [171], water erosion due to high-intensity rainfall events is the most
active exogenous geomorphic process.
In arid environments, water erosion is recognized as the main soil degradation
process. Indeed, in many arid lands, soil erosion is greater than might be expected
considering the associated low annual rainfall rate [166, 172]. For the Patagonian
rangelands, the rainfall erosivity, a measure of the rainfall potential to produce soil
erosion, was determined from the mean annual precipitation by Gaitan et al. [27],
and was among the lowest in the Argentina. Low vegetation cover, steep topogra-
phy, low infiltration capacity, and some high-intensity thunderstorms can be iden-
tified as erosion causative factors. Low soil infiltration rate due to low vegetation
cover and soil organic matter, and subsequent crust formation may produce runoff to
attain significant values in degraded rangelands [173]. Spatial variation in runoff and
erosion resulting from differences in soil surface conditions has been well
documented [169, 174]. Extensive woody plant encroachment and the consequent
decrease of grass cover, altered runoff and soil erosion across most of the drylands,
because the bare soil between shrubs is highly susceptible to water erosion
[175]. Although laminar erosion is the dominant process and affects extensive
areas, concentrated rill and gully erosion is common in lowland soils, principally
in wetlands (mallines) (Fig. 8).
There are different estimates of the area affected by water erosion in Patagonia. A
recent study [176] indicates that more than 9 Mha are affected by water and wind
erosion in the central and southern portion of the Patagonian rangelands (Chubut,
Santa Cruz, and Tierra del Fuego provinces). For the Chubut province, a total of
2.3 Mha are affected by water erosion [177]. Results obtained by mean of the USLE
model indicate that the Patagonian rangelands present erosion rates >10 t ha-1 year-
1
in almost 50% of its territory. Low vegetation covers and long and steep slope
account for these high values [27]. A previous work found high rates of soil erosion
in two areas of northeastern Patagonia, a pediment-like plateau and a flank pediment.
A dendrochronological analysis from exposed roots of a dwarf shrub in these areas
indicated erosion rates of 2.4 and 3.1 mm year-1, for the dominant soils, a Xeric
Haplocalcids and Xeric Calciargids, respectively [178]. These values were equiva-
lent to 28.8 and 38.4 mg ha-1 year-1. Soil erosion selectively reduces soil nutrients
and organic matter, thereby impacting land’s productive capacity [179, 180].
28 D. S. Fernández et al.

Fig. 8 Gully erosion in a wet meadow (mallin) of the central area of Patagonia. Gullies lower the
water table and accelerate the water circulation throughout these ecosystems favoring their desic-
cation. Photo by CM Rostagno, October 2015, El Escorial, Chubut

Wind erosion is an important process of land degradation in the arid and semi-arid
grazing lands of Patagonia. It occurs under strong winds over dry sandy low
vegetated soils (a common condition in the Patagonian rangelands). Wind erosion
and deposition as natural processes have affected different areas of the Patagonian
rangelands as the southern part of Peninsula Valdés where some dune fields origi-
nated in the west coast have been advancing since pre-settlement times
[181]. The supply area of the sediments forming these dune fields are the sandy
beaches of the Nuevo Gulf [182]. These loose, sand-sized sediments are transported
inland by the prevailing westerly winds. However, after the introduction of sheep in
this area at the beginning of the last century [183], sheep grazing has accelerated
wind erosion in some stabilized dunes under the form of blowouts [184]. Blowouts,
the most common aeolian erosional landforms in dune landscapes in arid zones, are
depressions or hollows formed by wind erosion on a preexisting sand deposit [185]
and are common features in vegetation-stabilized dune fields [186]. Other forms of
wind erosion are the erosion tongues, a highly visible wind erosion form in aerial
photos or satellite images that were originally described for the west area of the
Patagonian rangelands [187]. These forms develop as sand is transported and
deposited in narrow strips parallel to the wind direction, the “tongue” consisting of
an active front of dunes or mounds, a central area where transportation is maximum,
and the area of origin, generally a dry lagoon, where blowing prevails [188]. A
narrow strip along the parallel of 46°S, in the boundary between the Santa Cruz and
Chubut provinces, where strong west winds prevails, is one of the heaviest affected
areas by wind erosion in Patagonia (Fig. 9).
Agricultural Land Degradation in Argentina 29

Fig. 9 An erosion tongue (1) in the SW of the Chubut province. The erosion tongue originated in
the Coyote lagoon (2) has a length of 60 km and an average width of 600 m. The dune field in the
extreme of the tongue advances at a mean rate of 600 m year-1

Toward the west, all along the Patagonian Andean Region of Argentina, soils are
mainly developed from volcanic ashes. The sub-humid sector, the ecotone between
the forest and the steppe, has been deeply modified by different land uses. Histor-
ically, extensive sheep and cows grazing was practiced in this region. Recently, more
intensive land uses such as feed-lots and agriculture were incorporated. Erosion
studies developed in the ecotone of Chubut province, based on fallout radionuclides
(Caesium-137) showed that soil losses in the last 50 years varied between 22 and
33 mg ha-1 year-1 under different land uses [170]. Water erosion studies proved that
volcanic soils are highly erodible when they are bare, while soils without
non-crystalline aluminosilicates are the most erodible ones [167]. Furthermore,
soil degradation processes in rangeland ecotone may include changes in mineralogy
[180]. Overgrazing may expose soil to desiccation, allowing the non-crystalline
materials to evolve to halloysite-type crystalline minerals [189], reducing the poten-
tial of soils to stabilize organic matter and its resistance to erosion [167, 180].
In this zone, agroforestry practices help to reduce soil erosion [190] and might
improve carbon sequestration [191]. Afforestation measures in many degraded areas
developed as an alternative to extensive sheep grazing have been implemented in the
rangelands ecotone zone [192]. Afforestation decreases wind erosion (Fig. 10),
serving, at the same time, as carbon sinks, particularly when done with locally
adapted species. Pine plantations are also controversial, among other aspects, due
to their high invasive potential in fire affected areas. Pines could also increase fire
intensity and/or frequency creating a potential positive feedback between invasion
and wildfires [193].
Wildfires are a cause of desertification, because fires reduce vegetation and litter
cover, soil fertility, and increase runoff and soil erosion. In the Patagonian
rangelands, accidental fires are common in north east portion, the area that mainly
belongs to the Monte phytogeographical province. Wildfires usually occurred after
30 D. S. Fernández et al.

Fig. 10 (a) Pine plantation in the extra Andean, subhumid region of the Neuquen province on
strongly degraded soils covered by desert pavements. (b) Degraded shrub steppe of Mulinum
spinosum, an increaser species in overgrazed areas in the ecotone, west of the Chubut province.
Photos: (a) J Irisarri, 2019, NW Neuquén province. (b) CM Rostagno, December 2015, 15 km W of
Esquel, Chubut

Fig. 11 The picture in the left (February 3, 1994) shows a dust storm in an area that was burnt on
January 21, 1994. Note the small sand accumulation (nebka) in the leeward area of a mound with a
partially burn shrub. The picture in the right shows a set of pins used to assess soil erosion in the
mounds and in the intermound area. The mean soil layer removed from the top of the mounds
during the 12 months following the fire event was equal to 27 mm, equivalent to 130 mg ha-1.
Sediment and partially burned organic matter accumulated in the leeward area of the mound. Photo
by CM Rostagno, February 3, 1994, 20 km W of Puerto Madryn, Chubut

2 or 3 years of rainfall above the average. During these wet years, grass biomass
increases and fine fuel, provided manly by grasses, accumulates and the risk of high
severity wildfires increases. Wildfires generally occur in the middle of the summer,
when strong winds, high temperatures, and low humidity occur. Under these condi-
tions, wildfires may affect thousands of hectares in few hours. In the areas affected
by extreme summer wildfires, soil erosion, mainly wind erosion, may attain high
values in few days (Fig. 11).
Agricultural Land Degradation in Argentina 31

Table 9 Irrigated and salt-affected soils area in the different provinces of Patagonia Region
Irrigated land Area affected by soil Area affected by soil
Province area (ha) salinization (ha) salinization (%)
Neuquén 14,600 4,380 30.0
Río Negro 79,320 22,500 28.3
Chubut 26,050 17,749 68.1
Santa Cruz 500 – –
Tierra del – – –
Fuego
Total for the 120,470 44,629 37.04
region
ND no data
Data source: Sánchez et al. [51]

3.5.2 Soil Salinization in Irrigated Areas

Soil salinization is an important cause of soil degradation in the arid and semi-arid
areas under irrigation. However, salt-affected soils, both saline and sodic, may
develop both under dryland and irrigated conditions [38]. In Patagonia there is
about 120,470 ha of irrigated land; in this area, the extent of salt-affected soils is
~44,600 ha (Table 9). Salinity and sodicity adversely affect the physical and
chemical soil properties as well as biological soil functions and cause degradation
of soil and water resources negatively affecting crop production. It may result from
the water table elevation, dissolution of salt presented in the irrigated soil profile, or
by salts added in the applied water [53]. Other factors may contribute to salt
concentration in the soil profile, including insufficient drainage network, poor
maintenance of collectors and main drains, soil compaction, natural cycles of
seawater intrusion as is the case of the lower Chubut valley [181], fertilizer appli-
cation, and soil parent material. The link between Welsh colonization, the begin-
nings of agricultural production, and soil salinization in the lower Chubut valley has
been described in detail by Bergmann [194].

3.5.3 Other Economic Activities Triggering Soil Degradation


in Patagonia

Other economic activities like oil and gas pipeline construction have only a localized
impact on soil degradation [195]. Soil compaction in Patagonia is also mainly
associated with this industry. In a clear-cut strip of a gas-pipeline installation in
NE Patagonia, the highest bulk density (1.43 g cm-3) was recorded in the area where
the traffic of heavy machinery was more intense. The penetrometer resistance data
was also significantly higher in this area, where values higher than 1.00 MPa were
recorded. As a result of the increment in soil compaction, the infiltration rate in this
area decreased significantly [195]. In the soil beneath shrub clumps of the control
32 D. S. Fernández et al.

area, bulk density and penetrometer resistance were 1.03 g cm-3 and 0.5 MPa,
respectively.
Occasional spills of oil and its by-products have caused contamination with
polycyclic aromatic hydrocarbon [196, 197]. It is known that hydrocarbon pollutants
modify the physical, chemical, and biological properties of soils [198]. Certain
technologies to deal with degradation caused by oil activity have been developed
in Patagonia, including conservation tillage in decapitated areas and the selection of
native shrub species to recover the vegetation structure [196, 199, 200].
Bioremediation is considered as relatively cost-effective and environmentally
friendly technology for decontamination of polluted soil. The successful application
of bioremediation depends on the presence of appropriate hydrocarbon-degrading
microorganisms and the environmental conditions in situ [197]. In this sense,
climatic and edaphic conditions in Patagonia are restrictive; cold and fluctuating
temperatures, low moisture contents, low nutrient levels, and high soil pH are the
conditions which most likely limit hydrocarbon degradation [201]. To overcome
these limitations, autochthonous bioaugmentation, i.e. the introduction of indige-
nous microorganisms to the site to be decontaminated was proposed as a remediation
alternative. Indigenous soil aliphatic and aromatic hydrocarbons degrading bacteria
with the ability to tolerate the environmental stresses were isolated from Patagonia
contaminated soils, which could be considered for autochthonous bioaugmentation
strategies [197, 202].
Although pollution problems associated with metalliferous mines are recognized
worldwide, they have been scarcely addressed in Patagonia. A recent work found
high levels of heavy metals, such as Cu, Zn, As, and Pb, in the surface horizons of
soils near a polymetallic mine which was abandoned more than 60 years ago in the
Chubut province [203]. Besides, heavy metal pollutants were also recorded in the
soils of a Patagonian salt marsh (a protected coastal area) that has been exposed to
mining wastes deposited 40 years ago, showing that these wastes are still a source of
metal contamination [204, 205]. Concentrations of Cd (45 mg kg-1), Pb
(42,853 mg kg-1), Cu (24,505 mg kg-1), and Zn (28,686 mg kg-1) in the soils of
the waste pile exceeded guidelines for agricultural, residential, and industrial land
uses [205].
Thus, the advance of these extractive industries on Patagonian lands can increase
the risks of soil degradation and contamination, with serious consequences on the
environment, health, and the agricultural land use.

4 Concluding Remarks

Land degradation is a serious and widespread problem in Argentina. Approximately


40% of the Argentina area is affected by one or more soil degradation processes. The
severity and extent of the degradation are not homogeneous being controlled prin-
cipally by land use, climate, and relief. Water and wind erosion and soil salinization
are present in all regions of the country and constitute the main degradation in terms
Agricultural Land Degradation in Argentina 33

Fig. 12 Surface area affected by main land-degradation processes in the different regions of
Argentina

of total affected area. Patagonia and Central regions show the higher percentages of
degradation affected areas with 69.2% and 51.1%, respectively (Fig. 12). These
results agree with estimations of the report of Ministry of Environment and Sustain-
able Development of Argentina, that shows that 38.5% of the land area of the
country is degraded [206]. Regarding agricultural lands, that occupy approximately
57 Mha in Argentina, 61% would be affected by one or more degradation processes
[21]. The Central Region of Argentina is clearly the most affected followed by the
North-West Region where single or multiple land-degradation processes were
assessed in both regions.
In Argentina, the main drivers of land degradation are land-use/cover change
(from woodland and grazing land to cropland) and agricultural intensification. It is
particularly problematic in dry sub-humid environments (400–600 mm annual
precipitation) due to the confluence of instability climate conditions (alternating
periods of droughts and floods) and agriculture expansion. Deforestation and
overgrazing are the main causes of human-induced soil erosion while agriculture
intensification and vegetation change are the main drivers of nutrient depletion. The
increasing demand for food commodities and the rise of market prices, has put
pressure to increase the productivity of existing agricultural soil and the efficiency of
agricultural production [2]. As a consequence of the high demand, narrow crop
rotations or even monoculture, principally soybean, was adopted leading to soil
degradation of agricultural lands. These changes in agricultural production contrib-
uted to an accelerated negative increase in nutrient balances and soil compaction.
Humid and sub-humid Argentinean plains present loess derived soil with high
content of fine silt that tend to develop shallow compaction from years of continuous
no-tillage farming that use heavy machinery [207, 208]. This compaction has
34 D. S. Fernández et al.

reduced soil infiltration rates and contributed to runoff losses during periods of high
precipitation [209].
The climate change related drivers of land degradation are gradual changes of
temperature, precipitation, and wind, as well as changes of the distribution and
intensity of extreme events [210]. In the Atlantic coast of Argentina, rainfall has
shown a clear positive increasing trend throughout the twentieth century
[211]. Mean annual precipitation in the humid Pampa increased by 35% in the last
half of the twentieth century. This raise of humidity increases susceptibility of soils
to water erosion in this area. In dryland areas of Argentina, increased air temperature
and evapotranspiration and decreased precipitation amount, in interaction with
climate variability and human activities, have contributed to desertification.
According to the Ministry of Environment and Sustainable Development of Argen-
tina the effects of climate change in Patagonia and Cuyo regions are linked with
increased aridity, extreme heat waves, and wildfire hazards. In Northeast region,
climate change produce increased temperatures and more intense droughts that affect
ecosystems and communities. In this region, a study made on climate change impact
on crops production reflected a reduction in maize yields in future scenarios asso-
ciated with increased temperatures that shortened the crop cycle more than with
water stress [212]. Instead, water stress will continue to be an important constraint to
soybean yield in the context of global warming, but this effect is strongly affected by
rainfall regimes.
The total loss of ecosystem services due to land-use/cover change (LUCC),
wetlands degradation, and use of land degrading management practices on grazing
lands was about US$75 billion in the year 2007 [213]. This amount represented 16%
of the country’s gross domestic product (GDP) for that year. The magnitude of the
problem highlights the need for an adequate land use planning, incentives and
actions from local authorities and producers in order to protect high value biomes.
Land-degradation processes described in this chapter put in focus the need of
sustainable land management in Argentina. In agricultural lands the strategy should
be oriented to minimize soil erosion, create positive SOC and nitrogen budgets,
enhance crop diversification, and improve structural stability of soils. The manage-
ment of land farming systems should include novel and diversify crop sequences in
order to improve yields and resource use. Studies on crop sequences of Pampa
Region have shown that repeated cereal crops reduce their productivities, while well
balanced sequences with greater diversity resulted in the highest productivities of
cereal crops [28, 214]. Besides, the inclusion of some species as cover crops in the
simplified cropping systems improves water and nitrogen use efficiency, compared
to the long alternative fallow periods between summer crops [30]. The replacement
of natural areas of forests, woody and grasslands by annual crops must be carried out
within a land use scheme for rural areas, especially in regions where agriculture and
livestock are expanding faster like in Northwest region.
The state of situation related to agrochemicals in agricultural soils suggests that
some efforts have to be made to improve the biocide management decisions toward a
more efficient biocide profiles in oil-seed and cereal crops because of its large sown
area in Argentina. Different management practices and technological alternatives are
Agricultural Land Degradation in Argentina 35

available in order to minimize the impact of agrochemicals in the environment


[107]. The management strategy should include crop rotation, use of varieties with
genetic resistance, calibration of planting and harvest times, mechanical weed
control, biological control of pests and weeds, restricted and strategic use of
pesticides.

References

1. INDEC (2021) Censo Nacional Agropercuario 2018: resultados definitivos. Ciudad Autónoma
de Buenos Aires
2. Choumert J, Phélinas P (2015) Determinants of agricultural land values in Argentina. Ecol
Econ 110:134–140. https://doi.org/10.1016/j.ecolecon.2014.12.024
3. Viglizzo E, Jobbágy E (2010) Expansión de la frontera agropecuaria en Argentina y su
impacto ecológico-ambiental. INTA, Buenos Aires
4. SAGyP-CFA (1995) El Deterioro de las Tierras en la República Argentina. Alerta Amarillo,
Buenos Aires, p 287
5. Casas RR (2017) La degradación del suelo en la Argentina. In: Vazquez M (ed) Manejo y
conservación de suelos. Con especial énfasis en situaciones argentinas. Asociación Argentina
de la Ciencia del Suelo, Buenos Aires, p 386
6. Casas RR (1998) Causas y evidencias de la degradación de los suelos en la región
Pampeana. In: David Rockefeller Center for Latin American Studies (ed) Hacia esa agricultura
productiva y sostenible en la pampa. Orientación Gráfica Editora S.R.L., Buenos Aires
7. Paruelo JM, Guerschman JP, Verón SR (2005) Expansión agrícola y cambios en el uso del
suelo. Ciencia Hoy 15:14–23
8. Paruelo JM, Guerschman JP, Piñeiro G, Jobbágy EG, Verón SR, Baldi G, Baeza S (2006)
Cambios en el uso de la tierra en Argentina y Uruguay: Marcos conceptuales para su análisis.
Agrociencia 10(2):47–61
9. Villarino SH, Studdert GA, Baldassini P, Cendoya MG, Ciuffoli L, Mastrángelo M, Piñeiro G
(2017) Deforestation impacts on soil organic carbon stocks in the Semiarid Chaco Region,
Argentina. Sci Total Environ 575:1056–1065. https://doi.org/10.1016/j.scitotenv.2016.09.175
10. MAGyP (2020) Estimaciones Agrícolas. Ministerio de Agricultura Ganadería y Pesca de la
República Argentina. https://datos.magyp.gob.ar/dataset/estimaciones.agricolas
11. Wingeyer AB, Amado T, Pérez-Bidegain M, Studdert GA, Perdomo Varela CH, Garcia FO,
Karlen DL (2015) Soil quality impacts of current South American agricultural practices.
Sustainability 7:2213–2242. https://doi.org/10.3390/su7022213
12. FAOSTAT (2020) Food and agriculture data. Economic and Social Development Department;
FAO, Rome
13. Bianchi E, Szpak C (2017) Soybean prices, economic growth and poverty in Argentina and
Brazil. Background paper to the UNCTAD-FAO Commodities and Development Report 2017
Commodity Markets, Economic Growth and Development
14. Eswaran H, Lal R, Reich PF (2001) Land degradation: an overview. In: Bridges EM, Hannam
ID, Oldeman LR, Pening de Vries FW, Scherr SJ, Sompatpanit S (eds) Responses to land
degradation. Proceedings of the 2nd International Conference on Land Degradation and
Desertification, Khon Kaen, Thailand. Oxford Press, New Delhi
15. Olsson L, Barbosa H, Bhadwal S, Cowie A, Delusca K, Flores-Renteria D, Hermans K,
Jobbágy E, Kurz W, Li D, Sonwa D, Stringer L (2019) Land degradation. In: Shukla PR,
Skea J, Calvo Buendia E, Masson-Delmotte V, Pörtner H, Roberts DC, Zhai P, Slade R,
Connors S, Van Diemen R, Ferrat M, Haughey E, Luz S, Neogi S, Pathak M, Petzold J,
Portugal Pereira J, Vyas P, Huntley E, Kissick K, Belkacemi M, Malley J (eds) Climate change
and land: an IPCC special report on climate change, desertification, land degradation,
36 D. S. Fernández et al.

sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosys-
tems. In press
16. Lavado RS, Taboada MA (2009) The Argentinean Pampas: a key region with a negative
nutrient balance and soil degradation needs better nutrient management and conservation
programs to sustain its future viability as a world agroresource. J Soil Water Conserv 64(5):
150–153
17. Borrelli P, Robinson DA, Fleischer LR, Lugato E, Ballabio C, Alewell C, Meusburger K,
Modugno S, Schütt B, Ferro V, Bagarello V, Van Oost K, Montanarella L, Panagos P (2017)
An assessment of the global impact of 21st century land use change on soil erosion. Nat
Commun 8:2013. https://doi.org/10.1038/s41467-017-02142-7
18. Piquer-Rodríguez M, Butsic V, Gärtner P, Macchi L, Baumann M, Gavier Pizarro G, Volante
JN, Gasparri IN, Kuemmerle T (2018) Drivers of agricultural land-use change in the Argentine
Pampas and Chaco regions. Appl Geogr 91:111–122. https://doi.org/10.1016/j.apgeog.2018.
01.004
19. Osinaga NA, Álvarez CR, Taboada MA (2018) Effect of deforestation and subsequent land
use management on soil carbon stocks in the South American Chaco. Soil 4:251–257. https://
doi.org/10.5194/soil-4-251-2018
20. Irurtia CB, Maccarini GD (1993) Erosión de suelos en la República Argentina. In: FAO
(ed) Erosión de Suelos en América Latina, Chile. pp 1–12
21. Prâvâlie R, Patriche C, Borrelli P, Panagos P, Rosca B, Dumitrascu M, Nita I, Sâvulescu I,
Birsan M, Bandoc G (2021) Environ Res 194:110697. https://doi.org/10.1016/j.envres.2020.
110697
22. Volante JN, Collado AD, Ferreyra EB, López C, Navarro Rau MF, Pezzola A, Puentes MI
(2009) Monitoreo de la cobertura y el uso del suelo a partir de sensores remotos. Resultados
2006-2009. Programa Nacional de Ecorregiones, INTA
23. Adamo SB, Crews-Meyer KA (2006) Aridity and desertification: exploring environmental
hazards in Jáchal, Argentina. Appl Geogr 26:61–85
24. Mirzabaev A, Wu J, Evans J, García-Oliva F, Hussein I, Iqbal M, Kimutai J, Knowles T,
Meza F, Nedjraoui D, Tena F, Türkeş M, Vázquez R, Weltz M (2019) Desertification. In:
Shukla P, Skea J, Calvo Buendia E, Masson-Delmotte V, Pörtner H, Roberts D, Zhai P,
Slade R, Connors S, Van Diemen R, Ferrat M, Haughey E, Luz S, Neogi S, Pathak M,
Petzold J, Portugal Pereira J, Vyas P, Huntley E, Kissick K, Belkacemi M, Malley J (eds)
Climate change and land: an IPCC special report on climate change, desertification, land
degradation, sustainable land management, food security, and greenhouse gas fluxes in
terrestrial ecosystems. In press
25. Fernández OA, Gil ME, Distel RA (2009) The challenge of rangeland degradation in a
temperate semiarid region of Argentina: The Caldenal. Land Degrad Dev 20:431–440
26. Altieri MA (1999) Applying agroecology to enhance productivity of peasant farming systems
in Latin America. Environ Dev Sustain 1:197–217
27. Gaitán J, Navarro M, Tenti Vuegen L, Pizarro M, Carfagno P, Rigo S (2017) Estimación de la
pérdida de suelo por erosión hídrica en la República Argentina. Instituto de Suelos – Centro de
Investigación de Recursos Naturales (CIRN), Centro Nacional de Investigaciones
Agropecuarias – INTA, Ediciones INTA, Buenos Aires
28. D’Acunto L, Andrade JF, Poggio SL, Semmartin M (2018) Diversifying crop rotation
increased metabolic soil diversity and activity of the microbial community. Agric Ecosyst
Environ 257:159–164. https://doi.org/10.1016/j.agee.2018.02.011
29. Cabrera AL (1976) Regiones fitogeográficas argentinas. In: Kugler WF (ed) Enciclopedia
argentina de agricultura y jardinería, 2nd edn. Acme, Buenos Aires, Tomo 2, Fascículo 1:1–85
30. Restovich SB, Andriulo AE, Portela SI (2012) Introduction of cover crops in a maize-soybean
rotation of the Humid Pampas: effect on nitrogen and water dynamics. Field Crop Res 128:62–
70. https://doi.org/10.1016/j.fcr.2011.12.012
Agricultural Land Degradation in Argentina 37

31. Colazo JC, Carfagno P, Gvozdenovich J, Buschiazzo D (2019) Soil erosion. In: Rubio G,
Lavado RS, Pereyra FX (eds) The soils of Argentina. World Soils Book Series. Springer, pp
239–250
32. Borrelli P, Robinson DA, Fleischer LR, Lugato E, Ballabio C, Alewell C, Meusburger K,
Modugno S, Schütt B, Ferro V, Bagarello V, Van Oost K, Montanarella L, Panagos P (2017)
An assessment of the global impact of 21st century land use change on soil erosion. Nat
Commun 8:2013. https://doi.org/10.1038/s41467-017-02142-7
33. Colazo JC, Panebianco JE, Del Valle HF, Godagnone RE, Buschiazzo DE (2008) Erosión
eólica potencial de suelos de Argentina. Efecto de registros climáticos de distintos
periodos. In: Actas del XXI Congreso Argentino de la Ciencia del Suelo Semiárido de la
Asociación Argentina Ciencia del Suelo, 13 al 16 de mayo de 2008, Potrero de los Funes,
Argentina
34. Colazo JC, Buschiazzo DE (2010) Soil dry aggregate stability and wind erodible fraction in a
semiarid environment of Argentina. Geoderma 159:228–236
35. Buschiazzo DE, Panebianco JE, Colazo JC (2014) Erosión eólica y cambio climático en suelos
de Argentina. In: Pascale C, Zubillaga M, Taboada M (eds) Los suelos, la producción
agropecuaria y el cambio climático: avances en Argentina. Ministerio de Agricultura de la
Nación, Buenos Aires, pp 376–384
36. FAO (2015) Estudio del potencial de ampliación del riego en Argentina. UTF/ARG/017
Desarrollo Institucional para la Inversión. UCAR – PROSAP, Roma
37. Casas RR, Rossi MS (2011) Manejo de campos salinos. Ciencia y tecnología. Supercampo.
Año XVII, N° 196
38. Pla Sentís I (2015) Advances in the prognosis of soil sodicity under dryland and irrigated
conditions. Int Soil Water Conserv Resour 2:50–63
39. Gasparri NI, Grau HR, Gutierrez Angonese J (2013) Linkages between soybean and neotrop-
ical deforestation: coupling and transient decoupling dynamics in a multi-decadal analysis.
Glob Environ Chang 23(6):1605–1614
40. Volante JN, Mosciaro MJ, Gavier-Pizarro GI, Paruelo JM (2016) Agricultural expansion in the
Semiarid Chaco: poorly selective contagious advance. Land Use Policy 55:154–165
41. Proyecto de monitoreo de deforestación en el Chaco seco (2019). http://www.
monitoreodesmonte.com.ar. Accessed 30 Jun 2021
42. Ministerio de Ambiente y Desarrollo Sostenible (2020). https://www.argentina.gob.ar/
ambiente/accion/deforestacion. Accessed 22 Jun 2021
43. Grau HR, Gasparri NI, Aide TM (2005) Agriculture expansion and deforestation in seasonally
dry forests of north-west Argentina. Environ Conserv 32(2):140–148. https://doi.org/10.1017/
S0376892905002092
44. Volante JN, Alcaraz-Segura D, Mosciaro MJ, Viglizzo EF, Paruelo JM (2012) Ecosystem
functional changes associated with land clearing in NW Argentina. Agric Ecosyst Environ
154:12–22. https://doi.org/10.1016/j.agee.2011.08.012
45. Busnelli J, Neder L del V, Sayago JM (2006) Temporal dynamics of soil erosion and rainfall
erosivity as geoindicators of land degradation in Northwestern Argentina. Quat Int 158:147–
161. https://doi.org/10.1016/j.quaint.2006.05.019
46. Fernández DS, Lutz MA (2014) Carta de peligrosidad geológica “San Fernando del Valle de
Catamarca”, provincias de Catamarca, Santiago del Estero y Tucumán. Servicio Geológico
Minero Argentino, Buenos Aires
47. Sayago JM, Collantes M (2016) Geopedology and land degradation in North-West
Argentina. In: Joseph Alfred Zinck JA, Metternicht G, Bocco G, Del Valle H (eds)
Geopedology. Springer, pp 441–455
48. Osinaga R, Osinaga N, Arzeno JL (2015) Provincia de Salta. In: FESIC (ed) El Deterioro del
Suelo y del Ambiente de la Argentina. FESIC-PROSA, Buenos Aires, Tomo II (7), pp
225–234
38 D. S. Fernández et al.

49. Tejerina Díaz FG (2010) Determinación de la Erosión Hídrica Potencial de los Suelos de la
Alta Cuenca del Río Bermejo (Salta), implementado en Formato S.I.G. Universidad Nacional
de Salta
50. Yapur SM (2010) Determinación de la Erosión Hídrica Potencial de los Suelos de la Cuenca
Inferior del Río Bermejo, Salta-Jujuy. Universidad Nacional de Salta
51. Powlson DS, Whitmore AP, Goulding KW (2011) Soil carbon sequestration to mitigate
climate change: a critical re-examination to identify the true and the false. Eur J Soil Sci
62(1):42–55. https://doi.org/10.1111/j.1365-2389.2010.01342.x
52. Ripley SW, Krzic M, Bradfield GE, Bomke A (2010) Land-use impacts on selected soil
properties of the Yungas/Chaco transition forest of Jujuy province, northwestern Argentina: a
preliminary study. Can J Soil Sci 90:679–683. https://doi.org/10.4141/CJSS09101
53. Sánchez RM, Guerra LD, Scherger M (2016) Evaluación de las aéreas bajo riego afectadas por
salinidad y/o sodicidad en Argentina. INTA, Buenos Aires
54. Jobbágy EG, Giménez R, Marchesini V, Diaz Y, Jayawickreme D, Nosetto M (2020) Salt
accumulation and redistribution in the dry plains of Southern South America: lessons from
land use changes. In: Taleisnik E, Lavado RS (eds) Saline and alkaline soils in Latin America.
Springer, pp 51–70. https://doi.org/10.1007/978-3-030-52592-7_3
55. Marchesini VA, Giménez R, Nosetto MD, Jobbágy EG (2017) Ecohydrological transforma-
tion in the Dry Chaco and the risk of dryland salinity: following Australia’s footsteps?
Ecohydrology 10(4):1822
56. Amdan ML, Aragon R, Jobbagy EG, Volante JN, Paruelo JM (2013) Onset of deep drainage
and salt mobilization following forest clearing and cultivation in the Chaco plains (Argentina).
Water Resour Res 49:6601–6612. https://doi.org/10.1002/wrcr.20516
57. Giménez R, Mercau J, Nosetto M, Páez R, Jobbágy E (2016) The ecohydrological imprint of
deforestation in the semiarid Chaco: insights from the last forest remnants of a highly
cultivated landscape. Hydrol Process 30(15):2603–2616. https://doi.org/10.1002/hyp.10901
58. Puchulu ME (2010) Los materiales parentales de los suelos y su relación con el
comportamiento de las sales en el sudeste de la provincia de Tucumán. Tesis doctoral de la
Universidad Nacional de Tucumán, Argentina
59. Fernández DS, Puchulu ME (2006) Los suelos del sudeste tucumano y su relación con la
geomorfología e hidrología regional. In: Actas del Actas del XX Congreso Argentino de la
Ciencia del Suelo, Salta, Argentina, 19–22 septiembre 2006
60. Maccagno P, Navone S, Trebino H (2015) Evolución del grado de desertificación y su relación
con los aspectos socioeconómicos en la cuenca del río Santa María, Catamarca, Argentina.
Revista de Investigación Agropecuaria 41(3):298–308
61. Navone S, Espoz-Alsina C, Maggi A, Introcaso R (2002) Monitoreo de la desertificación en
los valles semiáridos del noroeste argentino: Desarrollo de un Sistema de Información
Geográfica empleando indicadores biofísicos y socieconómicos. Revista de Teledetección
18:5–19
62. Collantes M, González L (2012) Mecanismos del proceso de desertificación en el valle de
Santa María, provincia de Tucumán (Argentina). Acta geológica 24(1–2):108–122
63. Maccagno P, Trebino T (2017) Valorización económica, social y ambiental de la
desertificación en el Departamento Santa María, Provincia de Catamarca, Argentina. Revista
Iberoamericana de Economía Ecológica 27:85–101
64. Pereyra FX, Fernández DS (2019) North-western Argentina soils. In: Gerardo Rubio G,
Lavado RS, Pereyra FX (eds) The soils of Argentina. World Soils Book Series. Springer, pp
123–134
65. García CL, Teich I, Gonzalez Roglich M, Kindgard AF, Ravelo AC, Liniger H (2019) Land
degradation assessment in the Argentinean Puna: comparing expert knowledge with satellite-
derived information. Environ Sci Pol 91:70–80
66. Fernández Turiel JL, Aceñolaza P, Medina ME, Llorenz JF, Sardi F (2001) Assessment of a
smelter impact area using surface soils and plants. Environ Geochem Health 23:65–78. https://
doi.org/10.1023/A:1011071704610
Agricultural Land Degradation in Argentina 39

67. Puchulu ME, Fernández DS, Moreno C (2009) Evidencias de contaminación de suelos por
actividades industriales en el Departamento Lules, Tucumán. Serie Monográfica y Didáctica,
vol 48. Tucumán, Argentina
68. Villegas D (2018) Degradación de Suelos. In: Gozalvez MR (ed) Estudio Geoambiental San
Antonio de Los Cobres, Provincias de Jujuy y Salta, Argentina. Instituto de Geología y
Recursos Minerales SEGEMAR, Buenos Aires
69. Logan TJ (1990) Chemical degradation of soil. In: Lal R, Stewart BA (eds) Advances in soil
science, vol 11. Springer, New York, pp 187–221. https://doi.org/10.1007/978-1-4612-
3322-0_6
70. Martínez Riera N, Soria N, Feldman G, Riera N (2006) Niveles de Plombemia y otros
marcadores, en niños expuestos a una fundición de plomo en Lastenia, Tucumán, Argentina.
RETEL Revista de Toxicología en Línea 11:12–22
71. Kirschbaum A, Murray J, Arnosio M, Tonda R, Cacciabue L (2012) Pasivos ambientales
mineros en el noroeste de Argentina: Aspectos mineralógicos, geoquímicos y consecuencias
ambientales. Revista Mexicana de Ciencias Geológicas 29(1):248–264
72. Do Campo M, Valenzuela MF, Ferro L (2020) Dispersión de contaminantes a partir de
residuos mineros de una antigua planta de fundición (Ingenio Muschaca) ubicada en el Distrito
Choya (Andalgalá, Catamarca). Rev Asoc Geol Argent 77(2):220–229
73. Fehlenber V, Baumann M, Nestor IgnacioGasparri NI, Piquer-Rodriguez M, Gavier-Pizarrod-
G, Kuemmerle T (2017) The role of soybean production as an underlying driver of defores-
tation in the South American Chaco. Glob Environ Chang 45:24–34
74. Murgida AM, González MH, Tiessen H (2014) Rainfall trends, land use change and adaptation
in the Chaco salteño region of Argentina. Reg Environ Chang 14:1387–1394
75. Fernández R, Sosa D, Pahr N, Von Wallis A, Bárbaro S, Albarracin S (2015) Provincia de
Misiones. In: Casas R, Albarracín G (eds) El deterioro del suelo y del ambiente en la
Argentina. Fundación para la Educación la Ciencia y la Cultura (FECIC)
76. Zurita JJ, López A, Brest E, Rojas J, Goytía Y, Bianconi A (2010) Zonificación RIAN Chaco y
Formosa. https://inta.gob.ar/documentos/zonificacion-rian-de-chaco-formosa. Accessed
14 May 2021
77. Dirección de Suelos y Agua Rural de la Provincia del Chaco (2017) Degradación de suelos en
la Región Chaqueña. In: Taller Regional NEA, Plan Nacional de Suelos Agropecuarios,
Misiones, 4 May 2017
78. Ledesma L, Casas L, Zurita J (1992) Los Suelos del Departamento Chacabuco, Provincia del
Chaco. INTA, Buenos Aires
79. Baridón JE, Casas RR (2014) Quality indicators in subtropical soils of Formosa, Argentina:
changes for agriculturization process. Int Soil Water Conserv Res 2(4):13–24. https://doi.org/
10.1016/S2095-6339(15)30054-X
80. Prause J, Soler J (2001) Cambios producidos en un suelo bajo labranza conservacionista y
siembre directa de algodón en el Chaco, Argentina. Agricultura Técnica 61(4):1–7. https://doi.
org/10.4067/S0365-28072001000400016
81. López AE (2018) Estudio del impacto del cultivo de soja (Glycine max (L.) Merr.) sobre
indicadores edáficos y productivos en tierras desmontadas en el sector sur del Departamento
Almirante Brown, provincia del Chaco. Tesis magister en Manejo y Conservación de Recursos
Naturales de la Universidad Nacional de Rosario
82. Baridón E, Pellegrini A, Lanfranco J, Cattani V (2012) Variación de la fracción orgánica por
agriculturización en Alfisoles subtropicales de Argentina. CienciAgro, Journal de Ciencia y
Tecnología Agraria 2(3):371–378
83. Pérez-Brandan C, Meyer A, Meriles J, Huidobro J, Schloter M, Vargas-Gil S (2019) Relation-
ships between soil physicochemical properties and nitrogen fixing, nitrifying and denitrifying
under varying land-use practices in the northwest region of Argentina. Soil Water Res 14(1):
1–9. https://doi.org/10.17221/192/2017-SWR
40 D. S. Fernández et al.

84. Jayawickreme D, Santoni C, Kim J, Jobbáyi E, Jackson R (2011) Changes in hydrology and
salinity accompanying a century of agricultural conversion in Argentina. Ecol Appl 21(7):
2367–2379
85. FAO (2015) AQUASTAT Perfil de País - Argentina. Organización de las Naciones Unidas
para la Alimentación y la Agricultura, Roma
86. Hall AJ, Rebella CM, Ghersa CM, Culot JP (1992) Field-crop systems of the Pampas. In:
Pearson CJ (ed) Ecosystems of the world. Elsevier, Amsterdam, pp 413–450
87. Viglizzo EF, Carreño LV, Pereyra H, Ricard F, Clatt J, Pincén D (2010) Dinámica de la
frontera agropecuaria y cambio tecnológico. In: Viglizzo E, Jobbágy E (eds) Expansión de la
frontera agropecuaria en Argentina y su impacto ecológico-ambiental. INTA, Buenos Aires,
pp 9–30
88. Zárate MA (2003) Loess of southern South America. Quat Sci Rev 22(18–19):1987–2006.
https://doi.org/10.1016/S0277-3791(03)00165-3
89. INTA-SAGyP (1990) Atlas de suelos de la República Argentina. Secretaría de Agricultura,
Ganadería y Pesca, Buenos Aires, pp 1–731
90. Moscatelli GN (1991) Los suelos de la región pampeana. In: Barsky O (ed) El desarrollo
agropecuario pampeano. INDEC-INTA-IICA, Buenos Aires, pp 11–76
91. Panigatti JL (2010) Argentina 200 años, 200 suelos. Editorial INTA, Buenos Aires
92. Panigatti JL (2016) Aspectos de la erosión de suelos en Argentina II. Asociación Argentina de
la Ciencia del Suelo
93. Lavado RS (2016) Historias sobre la sustentabilidad de los sistemas de producción
agropecuaria y su vinculación con la fertilidad del suelo y el uso de fertilizantes. In: Lavado
RS (ed) Sustentabilidad de los agrosistemas y uso de fertilizantes. Orientación Gráfica, AACS,
Buenos Aires
94. Rabinovich JE, Torres F (2004) Caracterización de los síndromes de sostenibilidad del
desarrollo: el caso de Argentina. Volumen 38 de Serie Seminarios y conferencias. CEPAL,
Santiago de Chile
95. Viglizzo EF, Ricard MF, Taboada MA, Vázquez-Amábile G (2019) Reassessing the role of
grazing lands in carbon-balance estimations: meta-analysis and review. Sci Total Environ 661:
531–542
96. Viglizzo EF, Pordomingo A, Castro M, Lértora F (2002) La sustentabilidad ambiental de la
agricultura pampeana oportunidad o pesadilla? Ciencia Hoy 12(8):38–51
97. Senigagliesi C, Ferrari M (1993) Soil and crop responses to alternative tillage practices. In:
Buxton DR, Shibles R, Forsberg RA, Blad BL, Asay KH, Paulsen GM, Wílson RF (eds)
International crop science. Crop Science Society of America, Madison, pp 27–35
98. Pesce G, Scoponi L, Galantini JA, Durán R, Sánchez MA, De Batista M, Chimeno P,
Cordisco M, Olivers G, Merino L, Gzain M (2017) Valoración de la sustentabilidad de
sistemas de labranza: estudio de caso en el so bonaerense. III Jornadas Nacionales de Suelos
de Ambientes Semiáridos y del II Taller Nacional de Cartografía Digital, Bahía Blanca,
7–8 September 2017
99. Galantini JA, Iglesias JO, Maneiro C, Santiago L, Kleine C (2006) Sistemas de labranza en el
sudoeste bonaerense: efectos de largo plazo sobre las fracciones orgánicas y el espacio poroso
del suelo. Rev Invest Agrop (RIA-INTA) 35:15–30
100. Chagas CI, Santanatoglia OJ, Castiglioni MG, Massobrio MJ, Kraemer FB, Palacín EA, Bujan
A (2010) Land use and degradation in a basin belonging to the Rolling Pampa of Argentina:
tendencies and processes. In: Casanova M, Ruiz G, Zagal E (eds) Proceeding Contribution in
extenso. 16th International Soil Conservation Organization Congress, Soil Science Society of
Chile, pp 330–335
101. Echeverría NE (2014) Incidencia de disturbios antropogénicos sobre el escurrimiento y la
erosión hídrica de suelos del sur de la región semiárida argentina. Tesis Magister en Ciencias
Agrarias de la UNS
Agricultural Land Degradation in Argentina 41

102. Chagas CI, Kraemer FB (2018) Escurrimiento, erosión del suelo y contaminación de los
recursos hídricos superficiales por sedimentos asociados a la actividad agropecuaria extensiva:
algunos elementos para su análisis. Editorial Facultad de Agronomía, Ciudad de Buenos Aires
103. Colazo JC, Mendez M, Buschiazzo DE (2018) Medición de la erosión eólica. Análisis y
evaluación de propiedades físico hídrica de los suelos. Ediciones INTA, Centro Regional La
Pampa-San Luis
104. Casas RR, Damiano F (2019) Buenas prácticas de manejo y conservación del suelo y del agua
en la Argentina. In: Casas RR, Damiano F (eds) Manual de Buenas Prácticas de Conservación
del Suelo y del Agua en Áreas de Secano. Tomos I y II. Centro para la Promoción del Suelo y
del Agua -PROSA-, Fundación para la Educación, la Ciencia y la Cultura (FECIC), Buenos
Aires
105. Landriscini MR, Galantini JA, Duval ME, Iglesias JO (2020) Cambios en las fracciones
orgánicas resistentes en suelos de la región pampeana. Ciencia del Suelo 38(2):203–211
106. Rubio G, Taboada M (2013) Arbol de decisión para diagnosticar la capacidad productiva de
suelos de la región pampeana. Ciencia del Suelo 31(2):235–243
107. Viglizzo E, Frank F (2010) Erosión del suelo y contaminación del ambiente. In: Viglizzo E,
Jobbágy E (eds) Expansión de la frontera agropecuaria en Argentina y su impacto ecológico-
ambiental. Ediciones INTA, Buenos Aires
108. Novelli L, Caviglia O, Melchiori R (2011) Impact of soybean cropping frequency on soil
carbon storage in Mollisols and Vertisols. Geoderma 167–168:254–260. https://doi.org/10.
1016/j.geoderma.2011.09.015
109. Otondo J, Jacobo EJ, Taboada MA (2015) Mejora de propiedades físicas por el uso de especies
megatérmicas en un suelo sódico templado. Ciencia del suelo 33(1):119–130
110. Panebianco JE, Buschiazzo D (2007) Erosion predictions with the Wind Erosion Equation
(WEQ) using different climatic factors. Land Degrad Dev 19(1):36–44
111. Varela MF, Barraco M, Gili A, Taboada MA, Rubio G (2017) Biomass decomposition and
phosphorus release from residues of cover crops under no-tillage. Agron J 109(1):317–326.
https://doi.org/10.2134/agronj2016.03.0168
112. Giménez JE, Imbellone PA, Mormeneo ML (2018) Propiedades, clasificación y distribución
de horizontes endurecidos, con especial referencia a fragipanes. In: Alvarez C, Imbellone P
(eds) Compactaciones naturales y antrópicas en suelos argentinos, vol 1. Asociación Argentina
de la Ciencia del Suelo –AACS, Capítulo, pp 1–57
113. Casas RR, Puentes MI (2009) Expansión de la frontera agrícola en la región chaqueña:
impacto sobre la salud de los suelos. In: Morello J, Rodríguez A (eds) El Chaco sin bosques:
la Pampa o el desierto del futuro. Orientación Gráfica Editora, Buenos Aires
114. López FM, Duval M, Martínez JM, Galantini JA (2018) Propiedades físicas en suelos bajo
siembra directa del sudoeste bonaerense. In: Alvarez C, Imbellone P (eds) Compactaciones
naturales y antrópicas en suelos argentinos, vol 18. Asociación Argentina de la Ciencia del
Suelo –AACS, Capítulo, pp 534–547
115. Agostini M, Domínguez GF, Studdert GA, Tourn SN (2018) Impacto de diferentes prácticas
de manejo sobre algunas propiedades físicas de suelos del sudeste bonaerense. In: Alvarez C,
Imbellone P (eds) Compactaciones naturales y antrópicas en suelos argentinos, vol 16.
Asociación Argentina de la Ciencia del Suelo –AACS, Capítulo, pp 495–512
116. Álvarez CR, Fernández PL, Taboada MA, Consentino DJ (2018) Compactación en sistemas
agrícolas y mixtos de la pampa ondulada argentina. In: Alvarez C, Imbellone P (eds)
Compactaciones naturales y antrópicas en suelos argentinos, vol 14. Asociación Argentina
de la Ciencia del Suelo –AACS, Capítulo, pp 452–477
117. Sobral R, Vignale A, Alfieri A, Pecorari C (1993) Suelos overos del NO bonaerense.
Antecedentes, descripción, evaluación y conclusiones. Boletín de divulgación técnica N°
101. Estación Experimental Agropecuaria Pergamino. INTA
118. Cantú M, Becker A (1999) El impacto del uso intensivo de la tierra en áreas templadas del
centro de la República Argentina. I Conferencia Científica Internacional Medio Ambiente
Siglo 21, Cuba
42 D. S. Fernández et al.

119. Cisneros J, Degiovanni A, Gonzalez J, Cholaky C, Cantero A, Cantero J, Tassile J (2015)


Erosión y degradación de suelos. Provincia de Córdoba. In: Casas R, Albarracín G (eds)
Degradación de Tierras en la República Argentina. FECIC, Argentina, Tomo II: 87–100
120. MAGyP (2012) Estimaciones Agrícolas. Ministerio de Agricultura Ganadería y Pesca de la
República Argentina. https://datos.magyp.gob.ar/dataset/estimaciones.agricolas. Accessed
10 May 2021
121. Cholaky C, Bonadeo E (2018) Compactación de suelos en el centro-sur de Córdoba: causas,
consecuencias y manejo. In: Alvarez C, Imbellone P (eds) Compactaciones naturales y
antrópicas en suelos argentinos, vol 12. Asociación Argentina de la Ciencia del Suelo –
AACS, Capítulo, pp 375–414
122. Cantú MP, Becker AR, Bedano JC, Schiavo HF (2007) Evaluación de la calidad de suelos
mediante el uso de indicadores e índices. Ciencia del Suelo 25(2):173–178
123. Becker AR, Cantú MP, Schiavo HF (1998) Alteración de la Dinámica del agua y sales por la
presencia de Fragipanes. Evidencias Micromorfológicas. Proceedings 16 Congreso Mundial
de la Ciencia del Suelo, Francia. CD Symposium 30. 7 pp
124. Sasal MC, Wilson MG, Bedendo DJ, Schulz G (2015) Capítulo Provincia de Entre Ríos. In:
Casas R, Albarracín G (eds) El Deterioro del Suelo y del Ambiente en la Argentina. PROSA,
FECIC, INTA, pp 111–120
125. Imhoff S, Pilatti MA, Carrizo ME, Masola MJ, Marano RP, Felli O (2018) Compactación en
suelos del centro y norte de Santa Fe. In: Alvarez C, Imbellone P (eds) Compactaciones
naturales y antrópicas en suelos argentinos, vol 8. Asociación Argentina de la Ciencia del
Suelo –AACS, Capítulo, pp 153–179
126. Cerana J, Wilson MG, De Battista JJ, Noir J, Quintero C (2006) Estabilidad estructural de los
Vertisoles en un sistema arrocero regado con agua subterránea. Revista de Investigaciones
Agropecuarias RIA 35(1):87–106
127. Buschiazzo DE, Roberto Z, Colazo JC, Panebianco J (2015) Provincia de La Pampa. In:
Casas R, Albarracín G (eds) El deterioro del suelo y del ambiente en la Argentina. Prosa,
Buenos Aires, pp 141–154
128. Duval ME, Galantini JA, Martínez JM, López FM, Wall LG (2015) Evaluación de la calidad
física de los suelos de la región pampeana: efecto de las prácticas de manejo. Ciencias
Agronómicas - Revista XXV(33):033–043
129. Ciampitti IA, Piccone LI, Garcia FO, Rubio G (2011) Phosphorus budget and soil extractable
dynamics in field crop rotations in Mollisols. Soil Sci Soc Am J 75:131–142
130. García FO, San Juan MFG (2013) La nutrición de suelos y cultivos y el balance de nutrientes:
¿Como estamos? Inf Agron Hisp 9:2–7
131. Milesi Delaye LA, Irizar AB, Andriulo AE, Mary B (2013) Effect of continuous agriculture of
grassland soils of the Argentine Rolling Pampa on soil organic carbon and nitrogen. Appl
Environ Soil Sci 2013:487865. https://doi.org/10.1155/2013/487865
132. Taboada MA, Damiano F, Lavado RS (2017) Suelos afectados por sales e inundaciones en la
Pampa Deprimida y el oeste bonaerense. In: Taleisnik E, Lavado RS (eds) Ambientes salinos y
alcalinos de la Argentina. Orientación Gráfica, Universidad Católica de Córdoba, Ciudad
Autónoma de Buenos Aires, pp 55–88
133. Aparicio VC, Barbacone A, Costa JL (2014) Efectos de la calidad de agua para riego
complementario sobre algunas propiedades químicas edáficas. Ciencia del suelo 32(1):95–104
134. Suarez DL, Wood JD, Lesch SM (2006) Effect of SAR on water infiltration under a sequential
rain-irrigation management system. Agric Water Manag 86:150–164. https://doi.org/10.1016/
j.agwat.2006.07.010
135. Caprile AC, Aparicio VC, Portela SI, Sasal MC, Andriulo AE (2017) Drenaje y transporte
vertical de herbicidas en dos molisoles de la pampa ondulada Argentina. Ciencia del Suelo
35(1):147–159
136. Primost JE, Marino DJ, Aparicio VC, Costa JL, Carriquiriborde P (2017) Glyphosate and
AMPA, “pseudo-persistent” pollutants under realworld agricultural management practices in
Agricultural Land Degradation in Argentina 43

the Mesopotamic Pampas agroecosystem, Argentina. Environ Pollut 229:771–779. https://doi.


org/10.1016/j.envpol.2017.06.006
137. Lutri VF, Matteoda E, Blarasin M, Aparicio V, Maldonado L, Quinodoz F, Giuliano Albo J,
Cabrera A, Giacobone E (2019) Glyphosate transport modeling in soil and its relationship with
groundwater contamination in a rural area of the Pampean Plain of Córdoba, Argentina.
Geophys Res Abstr 21:EGU2019-1934
138. Pérez DJ, Iturburu FG, Calderon G, Oyesqui LA, Gerónimo ED, Aparicio VC (2021)
Ecological risk assessment of current-use pesticides and biocides in soils, sediments and
surface water of a mixed land-use basin of the Pampas region, Argentina. Chemosphere
263:128061. https://doi.org/10.1016/j.chemosphere.2020.128061
139. Alonso LL, Demetrio PM, Etchegoyen MA, Damián Marino J (2018) Glyphosate and atrazine
in rainfall and soils in agroproductive areas of the pampas region in Argentina. Sci Total
Environ 645:89–96. https://doi.org/10.1016/j.scitotenv.2018.07.134
140. Lupi L, Miglioranza KS, Aparicio VC, Marino D, Bedmar F, Wunderlin DA (2015) Occur-
rence of glyphosate and AMPA in an agricultural watershed from the southeastern region of
Argentina. Sci Total Environ 536:687–694. https://doi.org/10.1016/j.scitotenv.2015.07.090
141. Aparicio VC, De Gerónimo E, Marino D, Primost J, Carriquiriborde P, Costa JL (2013)
Environmental fate of glyphosate and aminomethylphosphonic acid in surface waters and soil
of agricultural basins. Chemosphere 9:1866–1873. https://doi.org/10.1016/j.chemosphere.
2013.06.041
142. Bento CP, van der Hoevena S, Yang X, Riksen MM, Mol HG, Ritsema CJ, Geissen V (2019)
Dynamics of glyphosate and AMPA in the soil surface layer of glyphosate-resistant crop
cultivations in the loess Pampas of Argentina. Environ Pollut 244:323–331. https://doi.org/10.
1016/j.envpol.2018.10.046
143. Mendez MJ, Aimar SB, Aparicio VC, Ramirez Haberkon NB, Buschiazzo DE, De
Gerónimo E, Costa JL (2017) Glyphosate and Aminomethylphosphonic acid (AMPA) con-
tents in the respirable dust emitted by an agricultural soil of the central semiarid region of
Argentina. Aeolian Res 29:23–29. https://doi.org/10.1016/j.chemosphere.2013.06.041
144. Okada E, Pérez D, De Gerónimo E, Aparicio V, Massone H, Costa JL (2018) Non-point source
pollution of glyphosate and AMPA in a rural basin from the southeast Pampas, Argentina.
Environ Sci Pollut Res 25:15120–15132. https://doi.org/10.1007/s11356-018-1734-7
145. UNEP (1997) World atlas of desertification.2nd edn. UNEP, London
146. Bianchi AR, Cravero SA (2010) Atlas climático digital de la República Argentina. Ediciones
Instituto Nacional de Tecnología Agropecuaria, Salta
147. Dascal G (2012) La vulnerabilidad de las tierras desertificadas frente a escenarios de cambio
climático en América Latina y el Caribe. Documentos de Proyectos 496, Naciones Unidas
Comisión Económica para América Latina y el Caribe (CEPAL)
148. ONDTyD (Observatorio Nacional de la Degradación de Tierras y Desertificación) (2019)
Therburg A, Corso ML, Stamati M, Bottero C, Lizana P, Pietragalla V (eds), Síntesis de
Resultados de la Evaluación de la Degradación de Tierras: 2012-2017. Mendoza, Argentina
149. FAO (2014) The water-energy-food nexus: a new approach in support of food security and
sustainable agriculture. Rome, FAO
150. Zubiate P, d’Hiriart A (2006) Hoja San Luis. Provincia de San Luis. INTA, San Luis
151. Guida-Johnson B, Abraham EM, Cony MA (2017) Salinización del suelo en tierras secas
irrigadas: perspectivas de restauración en Cuyo, Argentina. Revista de la Facultad de Ciencias
Agrarias de la Universidad Nacional de Cuyo 49(1):205–215
152. CNA (Censo Nacional Agropecuario) (2002) Resultados del Censo Nacional Agropecuario.
INDEC, Buenos Aires
153. Morábito J, Mirábile C, Pizzuolo P, Tozzi D, Manzanera M, Mastrantonio L (2004) Salinidad
de suelos regadíos e incultos en el oasis norte de Mendoza-Argentina. Actas del XIX Congreso
Argentino de la Ciencia del Suelo, Paraná, Entre Ríos-Argentina
44 D. S. Fernández et al.

154. Morábito J, Mirábile C, Manzanera M, Cappé O, Tozzi D, Mastrantonio L (2010) Evolución


de la salinidad de suelos regadíos e incultos en el área del Río Mendoza. XX congreso nacional
del agua—III simposio de recursos hídricos del cono sur. Conagua 2005. Argentina
155. Tozzi F, Mariani A, Vallone R, Morábito J (2017) Evolución de la salinidad de los suelos
regadíos del río Tunuyán Inferior (Mendoza - Argentina). Rev FCA UNCUYO 49(1):79–93
156. Vallone RC, Maffei JA, Morábito J, Mastrantonio L, Lipinski V, Filippini M, Olmedo F,
Castro G (2007) Mapa utilitario de suelos y riesgo de contaminación edáfica en los oasis
irrigados de la provincia de Mendoza. Jornadas de Investigación en Recursos Hídricos,
Espacio para la Ciencia y Tecnología, Mendoza
157. Vallone RC, Martínez LE, Olmedo FG, Sari SE (2021) Effects of salinity on vineyards and
wines from Mendoza, Argentina. In: Taleisnik E, Lavado RS (eds) Saline and alkaline soils in
Latin America. Springer Nature Switzerland, pp 161–176
158. Vallone RC, Martínez LE, Olmedo GF, Sari SE (2017) Influencia de la salinidad edáfica en el
contenido de cloruros y potasio de mostos y vinos en la cuenca del río Mendoza, Argentina. In:
Actas de la V Reunión de la Red Argentina de salinidad, villa mercedes, San Luis
159. Santoni CS, Jobbágy EG, Marchesini V, Contreras S (2008) Diferentes usos del suelo:
consecuencias sobre balance hídrico y drenaje profundo en zonas semiáridas. Actas del XXI
Congreso Argentino de la Ciencia del Suelo, AACS, San Luis
160. Jobbágy EG, Nosetto MD, Santoni CS, Baldi G (2008) El desafío ecohidrológico de las
transiciones entre sistemas leñosos y herbáceos en la llanura Chaco-Pampeana. Ecol Austral
18:305–322
161. Prohaska F (1976) The climate of Argentina, Paraguay and Uruguay. In: Schwerdtfeger E
(ed) Climate of Central and South America. World survey of climatology. Elsevier, Amster-
dam, pp 57–69
162. Von Thüngen J, Lanari MR (2010) Profitability of sheep farming and wildlife management in
Patagonia. Pastoralism 1:274–290
163. Huerta G (1991) Análisis económico de las explotaciones ovinas: Región Patagonia Norte. In:
FECOLAN (ed) Análisis económico de las explotaciones ovinas. 7a Jornadas Cooperativas de
Lanas, Buenos Aires, 23–24 August 1991, pp 99–106
164. Ares JO, Beeskow AM, Bertiller MB, Rostagno CM, Irisarri MP, Anchorena J, Defossé GE,
Merino C (1990) Structural and dynamic characteristics of overgrazed grasslands of northern
Patagonia, Argentina. In: Breymeyer A (ed) Managed grasslands. Elsevier, Amsterdam, pp
149–175
165. Del Valle HF, Elissalde NO, Gagliardini DA, Milovich J (1998) Status of desertification in the
Patagonian region: assessment and mapping from satellite imagery. Arid Soil Res Rehabil
12(2):1–27
166. Rostagno C, Coronato F, Del Valle H, Puebla D (1999) Runoff and erosion in five land units of
a closed basin of Northeastern Patagonia. Arid Soil Res Rehabil 13:281–292
167. La Manna L, Buduba C, Rostagno CM (2016) Soil erodibility and quality of volcanic soils as
affected by pine plantations in degraded rangelands of NW Patagonia. Eur J For Res 135:643–
655
168. Rostagno CM, Degorgue G (2011) Desert pavements as indicators of soil erosion on aridic
soils in north-east Patagonia (Argentina). Geomorphology 134:224–231
169. Chartier M, Rostagno CM, Videla L (2013) Selective erosion of clay, organic carbon and total
nitrogen in grazed semiarid rangelands of northeastern Patagonia, Argentina. J Arid Environ
88:43–49
170. La Manna L, Gaspar L, Tarabini M, Quijano L, Navas A (2019) 137Cs inventories along a
climatic gradient in volcanic soils of Patagonia: potential use for assessing medium term
erosion processes. Catena 181:104089. https://doi.org/10.1016/j.catena.2019.104089
171. Del Valle HF, Blanco PD, Hardtke LA, Metternicht G, Bouza PJ, Bisigato A, Rostagno CM
(2016) Contribution of open access global SAR mosaics to soil survey programs at regional
level: a case study in North-Eastern Patagonia. In: Zinck JA, Metternicht G, Bocco G, Del
Valle HF (eds) Geopedology. Springer, pp 321–346
Agricultural Land Degradation in Argentina 45

172. Renard K (1980) Estimating erosion and sediment yield from rangelands. In: Proceedings of
the Symposium on Watershed Management. ASCE, Reston, pp 164–175
173. Rostagno CM (1989) Infiltration and sediment production as affected by soil surface condi-
tions in a scrubland of Patagonia, Argentina. J Range Manag 42:382–385
174. Yair A, Enzel Y (1997) The relationship between annual rainfall and sediment yield in arid and
semi-arid areas. The case of the northern Negev. Catena Suppl 10:121–135
175. Parizek B, Rostagno CM, Sottini R (2002) Soil erosion as affected by shrub encroachment in
northeastern Patagonia. J Range Manag 55:43–48
176. Oliva G, García G, Ferrante D, Massara V, Rimoldi P, Díaz B, Paredes P, Gaitán J (2017)
Estado de los Recursos Naturales Renovables en la Patagonia Sur Extra andina. INTA Centro
Regional Patagonia Sur, Trelew
177. Salomone J, Llanos ME, San Martín A, Elissalde N, Behr S (2008) Uso del suelo y
degradación de tierras en la provincia del Chubut: evolución en los últimos veinte años.
Semiárido: un desafío para la Ciencia del Suelo. Actas del XXI Congreso Argentino de la
Ciencia del Suelo, San Luis
178. Chartier MP, Rostagno CM, Roig FA (2009) Soil erosion rates in rangelands of northeastern
Patagonia: a dendrogeomorphological analysis using exposed shrub roots. Geomorphology
106:344–351
179. Chartier MP, Rostagno CM, Pazos GE (2011) Effects of soil degradation on infiltration rates in
grazed semiarid rangelands of northeastern Patagonia, Argentina. J Arid Environ 75:656–661
180. La Manna L, Gaspar L, Rostagno CM, Quijano L, Navas A (2018) Soil changes associated
with land use in volcanic soils of Patagonia developed on dynamic landscapes. Catena 166:
229–239
181. Del Valle HF, Blanco PD, Sione W, Rostagno CM, Elissalde NO (2008) Assessment of salt-
affected soils using multisensor radar data: a case study from Northeastern Patagonia
(Argentina). In: Metternicht G, Zinck A (eds) Remote sensing of soil salinization: impact on
land management. Taylor and Francis, pp 155–173
182. Haller MJ, Monti AJ, Meister CM (2000) Hoja Geológica 4363-I: Península Valdés, Provincia
del Chubut. Programa Nacional de Cartas Geológicas de la República Argentina, 1:250000.
Boletín 266. Servicio Geológico Minero Argentino, Buenos Aires
183. Defossé GE, Bertiller MB, Rostagno CM (1992) Range management and development in
Argentine Patagonia: past, present and perspectives for the future. In: Perrier GK, Gay CW
(eds) Proceedings of the International Rangeland Development Symposium, February 1992.
Society for Range Management, Spokane, pp 12–21
184. Blanco PD, Rostagno CM, Del Valle HF, Beeskow AM, Wiegand T (2008) Grazing impacts
in vegetated dune fields: predictions from spatial pattern analysis. Rangel Ecol Manag 61(2):
194–203
185. Hesp PA (2002) Foredunes and blowouts: initiation, geomorphology and dynamics. Geomor-
phology 48:245–268
186. Livingstone I, Warren A (1996) Aeolian geomorphology: an introduction. Addison Wesley
Longman, Essex
187. Monteith N (1972) Estudio sobre la erosión en la Patagonia. Informe final Proyecto INTA-
FAO para el desarrollo ovino en la Patagonia. EEA INTA, Bariloche
188. Castro JM (1983) Manual para la recuperación de áreas erosionadas en la Patagonia. INTA,
EEA Trelew
189. Parfitt RL, Wilson AD (1985) Estimation of allophane and halloysite in three sequences of
volcanic soils, New Zealand. Catena 7:1–8
190. Tarabini M, Gómez F, La Manna L (2019) Ceniza volcánica reciente como indicadora de
retención de partículas en suelos de la Patagonia Andina. Ciencia de Suelo 37:101–112
191. Buduba CG, Defossé GE, Irisarri JA (2017) Impact of ponderosa pine afforestations on soil
organic matter (SOM) in semiarid steppes of western Patagonia, Argentina. Agroforest Syst
91:895–900. https://doi.org/10.1007/s10457-016-9963-6
46 D. S. Fernández et al.

192. Irisarri J, Dufilho AC, Buduba C, Muguerza D (2019) Provincia de Neuquén. In: Casas R,
Damiano F (eds) Manual de buenas prácticas de conservación del suelo y del agua en áreas de
secano, tomo II. Fundación para la Educación, la Ciencia y la Cultura, Buenos Aires, pp
105–144
193. Raffaele E, Nuñez MA, Eneström J, Blackhall M (2016) Fire as mediator of pine invasion:
evidence from Patagonia, Argentina. Biol Invasions 18:597–601
194. Bergmann JF (1971) Soil Salinization and Welsh Settlement in Chubut, Argentina. Cahiers de
géographie du Québec 15:361–369
195. Kowaljow E, Rostagno CM (2008) Efectos de la instalación de un gasoducto sobre algunas
propiedades del suelo superficial y la cobertura vegetal en el NE de Chubut. Ciencia del Suelo
26:51–62
196. Barquín M, Ríos SM, Nudelman N (2011) Test de toxicidad sobre la especie Atriplex lampa,
su aplicación para la evaluación de riesgo asociado a derrames de petróleo en suelos de la
Patagonia. Rev Toxicol 28(2):135–139
197. Madueño L, Coppotelli BM, Alvarez HM, Morelli IS (2011) Isolation and characterization of
indigenous soil bacteria for bioaugmentation of PAH contaminated soil of semiarid Patagonia,
Argentina. Int Biodeterior Biodegrad 65:345–351
198. Martínez MVE, López FS (2001) Efecto de hidrocarburos en las propiedades físicas y
químicas de suelo arcilloso. TERRA 19:9–17
199. Molina Sánchez D, Luque J, Amari M, Mac Karthy R, Lisoni C, Quinteros J, Macheprang O
(1995) Recuperación edáfica y revegetación de áreas afectadas por derrames de petróleo en la
región árida fría de Patagonia. Segundo Simposio de Producción de Hidrocarburos. Argentina,
pp 597–601
200. Nakamatsu V, Ciano N, Luque J (1998) Adaptación de especies vegetales nativas plantadas en
suelos contaminados con hidrocarburos en Patagonia. 51 Annual Meeting. Society for Range
Management, Guadalajara, México
201. Haro PA, Perez JD, Alvarez AF, Silva RA, Alvarez HM (2007) Bioestimulation treatments of
hydrocarbon-contaminated soil in semiarid Patagonia, Argentina. In: Proceeding of the SPE
10th Latin American and Caribbean Petroleum Engineering Conference 2, pp 693–698
202. Pucci OH, Bak MA, Peressutii SR, Kledp I, Hartig C, Álvarez HM, Wünsche L (2000)
Influence of crude oil contamination on the bacterial community of semiarid soils of Patagonia
(Argentina). Acta Biotechnol 20:129–146
203. Vicente FR, Valenzuela MF (2019) Efectos de una escombrera polimetálica sobre las
propiedades de los suelos aledaños, Lago Fontana, Chubut. Párrafos Geográficos 18(2):35–54
204. Idaszkin YL, Alvarez M, Carol E (2017) Geochemical processes controlling the distribution
and concentration of metals in soils from a Patagonian (Argentina) salt marsh affected by
mining residues. Sci Total Environ 596–597:230–235
205. Marinho C, Giarratano HE, Esteves JL, Narvarte MA, Gil MN (2017) Hazardous metal
pollution in a protected coastal area from Northern Patagonia (Argentina). Environ Sci Pollut
Res 24:6724–6735
206. MAyDS (Ministerio de Ambiente y Desarrollo Sostenible) (2020) Informe del Estado del
Ambiente 2019. Primera edición, volumen combinado, Ciudad Autónoma de Buenos Aires
207. Álvarez CR, Fernández PL, Taboada MA (2012) Relación de la inestabilidad estructural con el
manejo y propiedades de los suelos de la región pampeana. Ciencia del Suelo 30(2):173–178
208. Álvarez CR, Taboada MA, Perelman SB, Morrás HJ (2014) Topsoil structure in no-tilled soils
in the Rolling Pampas, Argentina. Soil Research 52:533–542
209. Lavado RS, Taboada MA (2009) The Argentinean Pampas: a key region with a negative
nutrient balance and soil degradation needs better nutrient management and conservation
programs to sustain its future viability as a world agroresource. J Soil Water Conserv 64(5):
150A–153A. https://doi.org/10.2489/jswc.64.5.150A
210. Lin KC, Hamburg SP, Wang L, Duh C, Huang CM, Chang CT, Lin TC (2017) Impacts of
increasing typhoons on the structure and function of a subtropical forest: reflections of a
changing climate. Nat Sci Rep 7:4911. https://doi.org/10.1038/s41598-017-05288-y
Agricultural Land Degradation in Argentina 47

211. IPCC (2007) Climate change: synthesis report. In: Pachauri RK, Reisinger A (eds) Contribu-
tion of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change. IPCC, Geneva
212. Casali L, Herrera JM, Rubio G (2021) Modeling maize and soybean responses to climatic
change and soil degradation in a region of South America. Agron J 113(2):1381–1393
213. Bouza ME, Aranda-Rickert A, Brizuela MM, Wilson MG, Sasal MC, Sione SM, Beghetto S,
Gabioud EA, Oszust JD, Bran DE, Velazco V, Gaitán JJ, Silenzi JC, Echeverría NE, De Lucia
MP, Iurman DE, Vanzolini JI, Castoldi FJ, Hormaeche JE, Johnson T, Meyer S, Nkonya E
(2016) Economics of land degradation in Argentina. In: Nkonya E, Mirzabaev A, Von Braun J
(eds) Economics of land degradation and improvement – a global assessment for sustainable
development. Springer Open, Switzerland. https://doi.org/10.1007/978-3-319-19168-3
214. Andrade JF, Poggiob SL, Ermácorac M, Satorrea EH (2017) Land use intensification in the
Rolling Pampa, Argentina: diversifying crop sequences to increase yields and resource use.
Eur J Agron 82:1–10. https://doi.org/10.1016/j.eja.2016.09.013
Agricultural Soil Degradation in Australia

Frederick A. Dadzie, Eleonora Egidi, Jana Stewart, David J. Eldridge,


Anika Molesworth, Brajesh K. Singh, and Miriam Muñoz-Rojas

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2 Agricultural Land Degradation Drivers and Impacts on Australian Soils . . . . . . . . . . . . . . . . . . . 53
2.1 Soil Erosion in Australia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.2 Soil Compaction in Australia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.3 Salinization and Acidification of Australian Soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.4 Soil Contamination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

F. A. Dadzie, J. Stewart, and D. J. Eldridge


Centre for Ecosystem Science, School of Biological, Earth and Environmental Sciences, UNSW
Sydney, Sydney, NSW, Australia
e-mail: [email protected]; [email protected]; [email protected]
E. Egidi
Hawkesbury Institute for the Environment, Western Sydney University, Penrith, NSW,
Australia
e-mail: [email protected]
A. Molesworth
ANU Institute for Climate, Energy & Disaster Solutions, Australia National University,
Canberra, ACT, Australia
B. K. Singh
Hawkesbury Institute for the Environment, Western Sydney University, Penrith, NSW,
Australia
Global Centre for Land Based Innovation, University of Western Sydney, South Penrith, NSW,
Australia
e-mail: [email protected]

M. Muñoz-Rojas (✉)
Centre for Ecosystem Science, School of Biological, Earth and Environmental Sciences, UNSW
Sydney, Sydney, NSW, Australia
Departamento de Biología Vegetal y Ecología, Facultad de Biología, Universidad de Sevilla,
Sevilla, Spain
e-mail: [email protected]

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 49
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 49–68, DOI 10.1007/698_2023_966,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023,
Published online: 25 March 2023
50 F. A. Dadzie et al.

3 Impacts of Climate Change on the Agricultural Sector in Australia and Implications


for Land Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4 Conclusions and Future Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

Abstract Australia is the fifth largest country by size, and one of the most arid in the
world. Agriculture is one of the key sectors in terms of spatial extent and economic
relevance. However, because of their ancient and weathered nature, Australian
agricultural soils are particularly susceptible to degradation processes such as ero-
sion, compaction, salinization, acidification, and contamination, which ultimately
lead to fertility loss, desertification, and thus decline in agricultural production and
food security. Agricultural systems are under further pressure due to the rapidly
intensifying threat of climate change. Consequently, protecting, maintaining, and
restoring healthy soils is a significant challenge and one of the key national priorities.
In this chapter, we review the major climatic, anthropogenic, and biotic drivers of
agricultural land degradation in Australia, illustrating the impacts of these processes
on agricultural production, discussing implications for soils and land sustainability
in a global change context and outlining potential mitigation strategies. We conclude
by highlighting current policies in place to resolve the challenges of agricultural land
degradation in the Australian continent, with emphasis on the importance of con-
certed and collaborative effort involving government, business, and civil society
sectors to address them effectively.

Keywords Desertification, Land degradation, Soil biodiversity loss, Soil erosion,


Soil health

1 Introduction

Australia is the fifth largest country by size and the largest island continent, covering
about 5% of the world’s land area. Given its large area there is a high diversity of
climate, vegetation, and soil, with eight of the 14 global ecoregions found within its
borders. About three-quarters of the continent is considered arid or semiarid with
highly variable rainfall, with the driest regions concentrated in the interior and west
of the continent [1].
Following a similar distribution, Tenosols, Kandosols, Rudosols, Sodosols, and
Vertisols comprise the majority of Australian soil types [2] (Fig. 1). Less than 6% of
Australia’s land mass is arable, with flat, barren, and heavily eroded landscapes that
are largely unproductive for agriculture [3]. This is mainly because Australian soils
are ancient and highly weathered, largely nutrient-poor, and low in fertility. Most of
the continent’s soil is characterized by low pH, total and available Phosphorus (P),
and total Nitrogen (N), and greater C:N and C:P ratios than other regions of the globe
[4]. The annual value of ecosystem services contributed by soil in Australia is
Agricultural Soil Degradation in Australia 51

Australian Soil Classification Map

Vertosol
Sodosol
Dermosol
Chromosol
Ferrosol
Kurosol
Tenosol
Kandosol
Hydrosol
Podosol
Rudosol
Calcarasol
Organosol
Anthroposol

Fig. 1 Australian Soil Classification. Based on the Australian Soil Classification Map. Version
1.0.0. Terrestrial Ecosystem Research Network (TERN). (Dataset) by Searle (2021). Used under
CC 4.0

estimated at Aust $570 billion [5]. Considering agricultural production alone, soil
contributes Aust $63 billion per year to the Australian economy [6].
In terms of area covered, the dominant land use in Australia is livestock grazing
of native vegetation (44.9%) followed by grazing of modified pastures (9.2%) per
cent (Fig. 2a). Nature conservation and other forms of protection cover 38.2% of
Australia’s land area, with a 16% (a total of 124.7 million has) occupied by forests
and woodlands (Fig. 2b).
Australia has long been heavily reliant on primary production, so livestock
grazing has been an integral component of the Australian agricultural landscape
since the arrival of Europeans. At a continental level, livestock grazing can broadly
be divided into three main categories: (1) highly industrialized systems where
animals are restricted to feedlot situations, (2) mixed livestock-cropping systems,
and (3) extensive pastoralism on unfertilized native pastures. Industrial feedlot
grazing is highly localized, and the environmental effects are limited to small-scale
nutrient production from urine and dung, waterlogging, and sedimentation
[7]. Mixed cropping-grazing systems are typically located in more mesic areas,
and soil-related environmental effects are generally related to soil erosion (mostly
rill and gully erosion), soil structural decline, loss of soil carbon (C), changes in
nutrient pools, and acidification [8]. The most extensive form of livestock grazing
occurs in areas of low pastoral productivity (rangelands) and occurs over about 70%
of the national land area. The gross value of Australian agriculture in 2018–2019 was
Aust $62.2 billion, with the top three agricultural commodities produced (ranked by
export value) being: cattle and calves at Aust $9.5 billion, wool at Aust $4.2 billion,
and wheat at Aust $3.7 billion [9]. In total, about 44% of the terrestrial land area is
52

Fig. 2 Australian land use at the national scale, 2010–2011. Source: Australian Bureau of Agricultural and Resource Economics and Sciences, Land Use of
Australia 2010–11, used under CC BY 3.0 (a), and extent of all forms of vegetation across Australia. Source: Australian Bureau of Agricultural and Resource
Economics and Sciences, Integrated Vegetation Cover dataset 2009, used under CC BY 2.5 (b)
F. A. Dadzie et al.
Agricultural Soil Degradation in Australia 53

dedicated to livestock grazing, and the annual value of livestock grazing for the
period 2019–2020 was about Aust $28 billion [10].
Australia’s soils are threatened by land degradation, defined as a reduction in land
condition through inappropriate land management coupled with effects of climate
[1, 11]. About a third of Australian soil is degraded and threatened with fertility loss,
yet soils are under increasing pressure due to increasing national and international
demand for food. Consequently, in a country where soils are inherently infertile by
world standards, protecting, maintaining, and restoring soils are significant chal-
lenges and one of the key national priorities [11]. In this chapter, we outline the main
drivers and impacts of agricultural land degradation and discuss their implications on
soils and land in a global change context.

2 Agricultural Land Degradation Drivers and Impacts on


Australian Soils

Agriculture is the dominant sector driving degradation, yet one of the sectors most
susceptible to the effects of deteriorating land condition. In Australia, the main
impacts of land degradation are erosion, surface compaction, salinization, acidifica-
tion, and contamination [12]. The drivers of this soil degradation are numerous and
include climate, human-induced disturbance, and biotic factors (Fig. 3). These
drivers generate pressures on the land that, if not properly managed, can result in
reduced fertility, and in arid areas, desertification. Fertility decline mainly involves
changes in soil biological (e.g., shifts in microbial community composition, biodi-
versity loss) and physicochemical (e.g., losses in C, nutrients, moisture, structure,
decrease in pH, increase in salinity) properties, and a reduction in the capacity of the
soil to support plant growth and productivity [13].

2.1 Soil Erosion in Australia

Soil erosion is associated with soil surface destabilization, nutrient depletion, altered
soil microbial community structure, and a reduced capacity of the soil to cycle
nutrients, decompose organic matter, and support plant production). An inherently
variable climate, combined with sparse vegetation cover over about much of the
continent, and the typically poorly-structured, shallow and infertile topsoil make the
Australian continent particularly susceptible to erosion [3, 14]. Extreme climatic
events, such as droughts, floods, and wildfires intensify the removal of soil by wind
and water, and these events are expected to intensify over the next century due to
changing climates [15], strengthening the negative effects of erosion on soil fertility.
The greatest rates of erosion occurred during the early years of European colonisa-
tion in Australia [16], with overgrazing, land clearing (Fig. 4a), and other land use
54

Fig. 3 Schematic representation of how climatic, anthropogenic, and biotic factors drive fertility loss in Australia
F. A. Dadzie et al.
Agricultural Soil Degradation in Australia 55

Fig. 4 Land clearing area


(left side), next to a native
vegetation remnant (right
side) in the Brigalow Belt
South bioregion (new South
Wales). Credit: Jana Stewart

changes such as conversion of grazing land to cropping contributing most to soil


erosion (Fig. 5a).
Cultivation can accelerate erosion rates above natural levels, which are thought to
be extremely low and effectively zero. Innovative practices embodied in conserva-
tion agriculture [13] can help to reduce erosion while maintaining yield and improv-
ing soil health.
High levels of grazing in the early 1800s led to extensive erosion and land
degradation, with devastating effects on soils and landscape processes
[14]. Today, most livestock are concentrated in the wheat-sheep belt in eastern
Australia, but considerable grazing still occurs on unimproved pastures (rangelands),
which make up about 70% of the land area. A national review of the impacts of
grazing on Australia’s terrestrial ecosystems [17] revealed that grazing reduced
ecosystem functions, which comprises measures of soil and plant health, by about
24%. Grazing also resulted in significant reductions in plant and litter cover, which
are linked to increased susceptibility to erosion. These negative effects were more
pronounced in semiarid and arid environments [17]. In drier arid, semiarid, and dry
subhumid areas, trampling by livestock can lead to compaction and destruction of
the thin biotic covering on the soil surface [18]. This has many negative effects,
56 F. A. Dadzie et al.

Fig. 5 Extensive soil erosion caused by (a) sheep grazing, and exacerbated by drought, in northern
Victoria, (b) rabbit burrowing in western New South Wales, and (c) horse trampling in the
Australian alps. Credit: David Eldridge

predisposing the soil to erosion [19] and reducing its capacity to mineralize C, N,
and P, which are critical elements for plant growth. Trampling also reduces habitat
for soil biota (fungi, bacteria, invertebrates) that are critical for the breakdown of
litter. Removal of the biocrust, which is most common in rangelands in winter-
dominant rainfall areas in southern Australia, can lead to declines in infiltration and
may increase runoff [18]. In northern Australia, cyanobacterial biocrusts are known
to fix up to 5 kg ha-1 of nitrogen annually and are critical to support the N balance of
summer-growing grasslands [20]. Overall, overgrazing reduces this N and leads to
declines in pastoral production, reduced resilience to drought, and increased risk of
erosion.
In addition to livestock overgrazing, introduced alien species are also responsible
for soil erosion across the Australian continent. For example, European rabbits
(Oryctolagus cuniculus), invasive herbivores that occur over much of the continent,
are widely recognized as contributing to soil erosion, nutrient decline, and pasture
deterioration due to their warren building activities [21] (Fig. 5b). Apart from their
effects on plant community composition and wind erosion, rabbit activity can result
in the transportation of P from the subsoil, enhancing the growth of exotic plants.
Unlike European livestock, rabbits do not use defined pathways or tracks, which
become conduits for water, and potentially link areas of bare soil and therefore
produce runoff and overland flow. The effects of kangaroos (Macropus spp.),
however, are more nuanced and controversial. High densities of kangaroos have
Agricultural Soil Degradation in Australia 57

been shown to lead to declines in plant cover [22], particularly during droughts [23],
but these effects are not obvious at low animal densities [24]. Finally, land degra-
dation can also result from unrestrained grazing by introduced feral herbivores. In
northern Australia, unrestricted grazing by feral horses, buffaloes, wild cattle,
camels, and donkeys can result in widespread erosion, leading to land degradation,
loss of perennial pastures, and fouling of watering points. Australia’s alpine grass-
lands have also become extremely overgrazed by feral horses [25] (Fig. 5c). Their
impacts not only damage fragile, humous rich soils, but also reduce the quality of
drinking water available for many communities in the Murray-Darling Basin.
Improved land management strategies have now been widely adopted over much
of the wheat-sheep belt in eastern Australia in order to reduce erosion and improve
crop yields. These include reduced or no-till farming, conservation farming, and the
use of crop rotations, which maintain a protective mulch on the surface to reduce
raindrop impact, water and sediment loss, and wind erosion [13]. The widespread
adoption of these control measures has reduced soil erosion over the past 30 years,
particularly in the southeastern states [26]. However, the effectiveness of these long-
term management practices varies markedly across the country, and the legacy effect
of past erosion events may be present for many years.
Approaches relating to grazing rely on the control of animal stocking duration
and intensity to increase plant growth and organic matter retention in soil [27]. For
example, in southern and eastern Australia, traditional forms of livestock manage-
ment such as continuous grazing have been replaced by time-controlled or rotational
grazing where different paddocks are allowed to rest on a rotational basis. Rotational
grazing systems are considered to be more effective at mitigating the effects of soil
erosion and have the potential to enhance soil organic C stocks, increase ecosystem
sustainability, and enhance production outcomes [28]. They are also likely to have
an important role in increasing resilience to the impacts of climate change
[14, 29]. In drier areas, additional mitigation strategies include exclusion fencing
to reduce native and feral animal populations [28] and the use of low-risk or
conservative grazing, where the aim is to maintain critical levels of plant cover on
the soil surface by running fewer animals [30]. However, the effectiveness of these
different approaches to reduce soil fertility loss depends largely upon soil types,
locations, and other environmental factors (e.g., topography), further complicating
the management, sustainability, and resilience of agricultural land in Australia [12].

2.2 Soil Compaction in Australia

The early colonisers described Australian soils as soft and loose; very different from
our present-day soils. Much of the soil softening resulted from the soil disturbing
activities of abundant small mammals such as the Greater bilby (Macrotis lagotis)
and Burrowing bettong (Bettongia lesueur) that were once widespread but are now
restricted to small areas of the country [31]. European settlement was also associated
with the introduction of hard-footed ungulates (sheep, cattle, horses), that unlike
58 F. A. Dadzie et al.

resident soft-footed macropods, did not co-evolve with Australian soils. The intro-
duction of large numbers of livestock, coupled with traditional intensive cultivation
practices such as tillage and intensive cropping with short rotation periods has led to
extensive soil compaction [32].
Soil compaction leads to lower soil biodiversity and fertility levels, due to
reduced capacity to store and supply nutrients and water, creating a negative
feedback loop that makes restoration difficult. The annual loss in agricultural
production from soil compaction in Australia is estimated to exceed Aust $850
million [33]. While compaction is a widespread issue in Australia, it is difficult to
diagnose compared with other forms of land degradation such as salinization and
erosion because it lacks clear soil surface indicators [31]. A recent management tool
based on a Bayesian approach to assessing the qualitative measures of compaction
[34] allows farmers to make an informed decision about the likely crop yield losses if
they ignore the problems of compaction.
Deep ripping of soils has been promoted as a technique to restore compacted
soils, but this is often only a temporary solution, and its effectiveness depends on soil
characteristics such as texture [32], with soils compacting faster and often deeper
than the initial compaction prior to deep ripping (plow pan). Current research on
combatting soil compaction focusses largely on the positive benefits of controlled
traffic farming, which results in lower levels of soil compaction indicators [35]. The
benefits of controlled traffic farming depend on soil type, but despite its promotion as
an effective method since the 1990s, adoption has been low, at about 30% nationally.
Other techniques include crop rotation with deep, tap-rooted species such as canola,
maintaining ground cover, and mulching, to increase soil porosity, structure, and
organic matter content indicators [36]. Novel application of biochar has also been
explored recently and shows promising results, alleviating compaction pressure on
growth and reducing NO2 emissions indicators [37]. Rotational grazing has been
shown to reduce soil compaction while enhancing soil health [28]. Given that
approximately half of Australia’s land use is dedicated to grazing [17], the imple-
mentation of even modest management changes would have long-term benefits on
soil porosity, structure, and organic matter content.

2.3 Salinization and Acidification of Australian Soils

The sustainable management of Australian soils requires a clear understanding of the


processes and mechanisms driving salinization and sodicity in Australia. Salt has
accumulated naturally in Australian soils over millennia. Most of this salt has a
marine origin, being deposited on the surface through aeolian activity during past
periods of aridity, leached through the soil profile, remobilized by rising ground
water, and concentrated on the surface through evaporation [1]. Agriculture in
Australia, however, is highly reliant on irrigation from surface or groundwater, but
these aquifers are often saline, with levels ranging from 15 to 150 dS m-1. These
levels are likely to increase as climate change reduces the availability of water and
Agricultural Soil Degradation in Australia 59

salinity concentrations increase [38]. Thus, poor soil and water management can
exacerbate the problems of salinization [38, 39]. The financial loss in agriculture
from salinity and sodicity in Australia is estimated at Aust $187 million and Aust
$1.03 billion, respectively [40].
Natural perennial native vegetation has historically kept groundwater tables low
[41], but the expansion of shallow-rooted annual plants associated with land use
change means that plants are no longer able to intercept and absorb rising ground-
water, exacerbating dryland salinity. The increase in concentrated salt at the soil
surface leads to secondary salinity, which affects more than one million hectares of
agricultural land in Western Australia, with estimated annual losses of Aust $0.5
billion [42]. Secondary salinization causes osmotic stress and yield declines and can
reduce microbial activity and diversity [43], leading to reduced nutrient availability
[41]. Leaching and flushing salt beyond the rooting depth are effective methods to
prevent a build-up of salts in the soil profile, though excessive leaching and flushing
can reduce water and nutrient use efficiency [44]. Poor quality saline water when
used for irrigation can lead to salt accumulation, particularly in drylands, so in these
situations, micro-irrigation methods are preferred [45], though they are expensive to
implement. Saline soils can be amended with gypsum to improve plant growth, but
this does not restore the osmotic imbalance in the soil [46], so additional treatments
such as the incorporation of mulch or organic matter are needed. Deep subsoil
ripping and slotting with gypsum can be used to reduce the effects of sodic soils
and reduce compaction [47]. More recent, cost-effective methods to improve soils
include the use of cyanobacteria to improve plant yields [48], though this is seldom
practiced in Australia. Planting salt-tolerant species has also been used to ameliorate
sodic and saline soils, with species such as Eucalyptus globulus and Eucalyptus
camaldulensis shown to grow well in moderate to highly saline soils [49].
Poor management practices are also responsible for increased acidification of
Australian soils, with over-use and leaching of nitrogen-based inorganic fertilizers
among the major contributors [50]. During acidification, excess soluble aluminum is
released, restricting water and nutrient uptake by reducing root growth. Declining
pH can also reduce soil microbial activity and diversity, reducing the activity of
agriculturally important associations such as mycorrhizae and legume nodulation
[51]. Acidic soils are generally treated with lime and gypsum. Application rates vary
depending on the extent of acidification in the soil profile, with surface applications
taking up to two decades to improve acidic subsoils [52]. As with salinization, deep
ripping and the application of lime is an appropriate strategy, but only where the risk
of soil erosion is low [53]. Deep rippling may mix topsoils and subsoils and therefore
effect seedling emergence [52]. These include tree plantings (agroforestry) to
improve water use efficiency and reduce nitrate leaching and excessive use of
fertilizers [54], and the planting of acid-tolerant pasture species such as rye grass,
subterranean clover, and some native grasses [55].
60 F. A. Dadzie et al.

2.4 Soil Contamination

Fertilizers, pesticides, sewerage sludges, organic amendments, and composts can be


critical non-point sources of contamination of agricultural soils in Australia. For
example, pollution from cattle grazing and P and N in runoff from excessive grazing
close to waterways can increase siltation and electrical conductivity of the water,
lowering water quality for instream invertebrates and humans that depend on these
river systems [12]. The main soil contaminants identified in Australian soils include
heavy metals, i.e., cadmium (Cd), copper (Cu), lead (Pb), mercury (Hg), and zinc
(Zn), and organic pollutants, such as pesticides and polycyclic aromatic hydrocar-
bons (PAHs) [56]. Among these contaminants in Australian agricultural soils, Cd is
considered that of greatest concern as it has the greatest potential to transfer to food
crops [57]. The main source of Cd in Australian soils is phosphatic fertilizers (i.e.,
those high in P), which contain up to 50 mg Cd kg-1 [58]. Historically, Australian
phosphatic fertilizers had higher concentrations than other countries. While the
effects on soils have been similar to those in continents such as Europe because of
lower inputs per surface area and reduced atmospheric contamination of Australian
soils [59], the use of such fertilizers in Australian pastures has led to substantial
increases in soil Cd concentrations. In general, higher amount of Cd can affect the
quality of agricultural soils and lead to phytotoxicity, decline of soil microbial
processes, and transfer of toxic metals to humans via increased crop uptake or intake
by grazing livestock [59]. In sandy grazing soils of South Australia and Western
Australia, livestock grazing of plants containing Cd can result in accumulation of
this element in meat, which can potentially have a large impact on human
health [60].
Of emerging concern is the presence of plastic contaminants in soils [61]. These
can enter agricultural soils from several sources such as biosolid application and
plastic mulching [62]. Biosolids are a by-product of the waste management industry,
are treated and recycled through applications in agriculture and other sectors such as
forestry or land rehabilitation. In Australia, it has been estimated that three-quarters
of the 327,000 tons of biosolids produced in 2017 were recycled through agriculture
applications [63]. This quantity of biosolids may have added up to 241 tons of
microplastics to agricultural soils in Australia for that year [62]. Another study
projected that between 2,800 and 19,000 tons of microplastics per year are added
to Australian farmlands through the application of biosolids, similar to numbers
reported in other countries [64]. Nanoplastics (<1 μm in at least one dimension) can
pose a larger threat than microplastics in soils as they can be absorbed by plants
within their cells [65]. The accumulation of microplastics in soils can alter soil water
evaporation and promotes drying, resulting in a reduced plant performance [66]. For
example, a recent study found a negative effect on the vegetative and reproductive
growth of wheat [67]. Moreover, microplastics can have a detrimental impact on
beneficial living soil organisms that perform critical soil functions to sustain their
fertility, e.g., nematodes, collembolans, soil mammals [59], resulting in alterations in
the food consumption and mortality of these organisms. The role of earthworms in
Agricultural Soil Degradation in Australia 61

the biogenic transport of microplastics can be a pollution source for groundwater and
food webs [68]. In Australia, a recent study showed a 200% increase in the
bioaccumulation of two contaminants (perfluorooctane sulfonate, perfluorooctanoic
acid) in earthworms exposed to microplastic-contaminated soil, resulting in a sig-
nificant reduction in their reproduction [69].

3 Impacts of Climate Change on the Agricultural Sector in


Australia and Implications for Land Degradation

Australian soils, ecosystems, and their unique biodiversity are already under con-
siderable stress from multiple human-induced impacts, with climate change and land
use change being major threats [12]. Plant and animal species have limited oppor-
tunity to adapt to climate change due to the physical and biotic characteristics of the
Australian environment (i.e., topography, habitat fragmentation, and low capacity
for dispersal) and the restricted geographic ranges of species [70]. Even under
modest future warming scenario (~1°C), the potential for negative effects on
Australian ecosystems and species is significant. For example, evidence suggests
that shrub cover will decline, and bare ground will increase in eastern Australia [15]
over the next century. In some dryland regions, higher land temperature, greater
evapotranspiration, and decreased precipitation amount, in interaction with climate
variability and human activities, have already contributed to desertification [15, 71].
Agriculture production, climate change, and land degradation are closely
interconnected in Australia. Climate change poses a serious and present threat to
the viability of Australia’s agricultural sector, as well as regional food security
[72]. Higher temperatures reduced rainfall and snowfall, and increased frequency
and intensity of extreme weather events (i.e., droughts, floods, bushfires, and dust
storms) have reduced farm productivity and produce quality, and increased farm
expenditure in many regions (Fig. 6). With typically less than 30-day supply of
non-perishable food and less than 5-day supply of perishable food, a low level of
on-hand food reserves makes Australia extremely vulnerable to supply chain dis-
ruptions via extreme weather events [73, 74].
These impacts directly influence food insecurity, not only nationally, but also
with countries reliant on Australian farm produce, as well as having a negative
impact on the sector’s economic value, and natural landscapes. Changing climates
also make land less productive for growing crops and raising livestock [71, 75] and
increase the likelihood that new pests and diseases will emerge, further reducing the
quality of agricultural produce [76]. Food nutritional quality can also be reduced by
rising temperatures and potential increases in mycotoxins. Thus, the climate change
effects on food security ultimately have many direct and indirect consequences on
human health and wellbeing.
62 F. A. Dadzie et al.

Fig. 6 Effect of 2000 to 2019 climate conditions on average farm business profit. Source: Source:
ABARES farmpredict [74], used under CC BY 4.0

4 Conclusions and Future Perspectives

Agricultural land degradation imposes significant social, economic, and environ-


mental costs to the Australian community, and most forms of degradation, such as
soil erosion and salinity, are difficult to reverse. Achieving sustainable soil manage-
ment is therefore critical to sustain a growing population and the global demand for
food, particularly in the face of an increasingly variable and harsher climate. The
constrained nature of most of Australian agricultural soils and the adverse interaction
with climate have made it difficult to develop a sustainable approach to mitigate land
degradation across the country. Indeed, maintaining and improving soil health,
mitigating degradation, and protecting the nation’s natural capital require the devel-
opment of strategies and approaches that are tailored to Australia’s unique environ-
mental attributes, with a concerted and collaborative effort involving government,
business, and civil society sectors to address them effectively. However, the imple-
mentation of effective strategies and policies has been historically compounded by
several logistic limitations, including the size and complexity of a vast and highly
diverse country with a relatively small population and skill base, and the existence of
disparate acts, regulations, and agencies across the different states and territories,
making it difficult to develop a harmonious approach to land management. Addi-
tionally, major technical constraints include the lack of standardized methods for
mapping and monitoring soil condition in Australia and a paucity of long-term
research sites to support efficient monitoring. Key steps to achieve a sustainable
soil management have been recently identified and include: (1) the establishment of
coordinated programs of soil research, development, and extension; (2) the
Agricultural Soil Degradation in Australia 63

engagement with experienced and highly motivated specialists; (3) the development
of up-to-date soil maps and monitoring programs at resolutions relevant to farm
management; and (4) the design of technical solutions for erosion control, seques-
tering carbon, and sustainable farming. The integration between public sector and
private sector data streams is also considered of utmost importance for the successful
implementation of such strategies and programs.
To overcome some of these challenges, a National Soil Strategy has recently been
developed in collaboration with state and territory governments, the National Soils
Advocate and other major stakeholders in soil science and land management
[11]. The Strategy is a 20-year plan that sets out how Australia will value, manage,
and improve its soil and aims to include specific and measurable targets to underpin
the progress measures and ensure the success of the Strategy can be clearly mea-
sured, monitored, and communicated. The Strategy will be implemented through a
National Action Plan, which will set out 3- to 5-year milestones and actions, subject
to 5-yearly reviews to incorporate changing or emerging soil-related priorities. The
Plan will include a first investment (2022) of AUD$196.9 million in new funding
over 4 years to implement the National Soil Strategy and an Interim Commonwealth
Action Plan. The plan will include increasing soil knowledge through standardized
data collection, management and storage, allowing for more informed decisions
using reliable, up-to-date, accessible information. This nationally coordinated strat-
egy is a first step toward a “soil-centric” approach to land management that can only
benefit the sustainable management of Australia’s natural resources.
Nevertheless, efforts to address climate change impacts on Australian land and
soils are still needed. With the agricultural sector’s vulnerability to climate change,
and its strong linkages with food security, social stability, human health and
wellbeing, and viability of the land and soils, the seriousness of the challenge cannot
be overstated. Urgent, cohesive, and evidence-based strategies are needed to miti-
gate emissions, discourage poor land practices, and facilitate improved social and
environmental resilience. Commitment to tackling climate change must be supported
and advanced by political, business, and community leaders at national, regional,
and local levels. A systemic view of climate change impacts and opportunities for
improvement are needed, such as the transition of Australia’s energy from fossil fuel
dependence to renewable sources. Clean energy is intrinsically linked to the reduc-
tion of climate disrupting greenhouse gasses that are threatening producer liveli-
hoods and environmental health. The capture and storage of carbon in agricultural
landscapes offers opportunities not only to lessen dangerous greenhouse gas con-
centrations but also to improve natural resources (such as soil health via increased
carbon concentration) and economic opportunity from emerging carbon markets.
The allocation of funding for climate mitigation and adaptation strategies represents
an investment in Australian social and political stability, and the avoidance of
worsening environmental deterioration, including land degradation, and food inse-
curity. The impacts Australia, and the rest of the world, will experience from climate
change will be determined by the strategies employed and actions taken today.
64 F. A. Dadzie et al.

References

1. Metcalfe DJ, Bui EN (2017) Australia state of the environment 2016: land, independent report
to the Australian Government Minister for the Environment and Energy, Australian Govern-
ment Department of the Environment and Energy, Canberra. https://doi.org/10.4226/94/
58b6585f94911
2. Searle R (2021) Australian soil classification map. Version 1.0.0. Terrestrial Ecosystem
Research Network (TERN). (Dataset). https://doi.org/10.25901/edyr-wg85
3. Orians GH, Milewski AV (2007) Ecology of Australia: the effects of nutrient-poor soils and
intense fires. Biol Rev 82:393–423. https://doi.org/10.1111/j.1469-185X.2007.00017.x
4. Eldridge DJ, Maestre FT, Koen TB, Delgado-Baquerizo M (2018) Australian dryland soils are
acidic and nutrient-depleted, and have unique microbial communities compared with other
drylands. J Biogeogr 45:2803–2814. https://doi.org/10.1111/jbi.13456
5. Bennett JM, McBratney A, Field D, Kidd D, Stockmann U, Liddicoat C, Grover S (2019) Soil
security for Australia. Sustainability 11:3416. https://doi.org/10.3390/su11123416
6. Jackson T, Zammit K, Hatfield-Dodds S (2018) Snapshot of Australian agriculture. Australian
Bureau of Agricultural and Resource Economics and Sciences, Canberra
7. McGinn S, Flesch T, Chen D, Crenna B, Denmead O, Naylor T, Rowell D (2010) Coarse
particulate matter emissions from cattle feedlots in Australia. J Environ Qual 39:791–798.
https://doi.org/10.2134/jeq2009.0240
8. Holt J (1997) Grazing pressure and soil carbon, microbial biomass and enzyme activities in
semi-arid northeastern Australia. Appl Soil Ecol 5:143–149. https://doi.org/10.1016/S0929-
1393(96)00145-X
9. NFF (2017) Food, fibre and forestry facts: a summary of Australia’s agricultural sector. National
Farmers Federation, Kingston, ACT. https://nff.org.au/media-centre/farm-facts/
10. ABS (2018) Agricultural commodities, Australia. Australian Bureau of Statistics 71210 https://
www.abs.gov.au/AUSSTATS/[email protected]/DetailsPage/7121.02017-18?OpenDocument
11. DAWE (2021) National Soil Strategy. Department of Agriculture, Water and the Environment,
Canberra, Australia. https://www.agriculture.gov.au/sites/default/files/documents/awe.gov.au/
publications. Last access 11 Apr 2022
12. Gretton P, Salma U (1996) Land degradation in the Australian Agricultural Industry. Industry
Commission, Canberra
13. Bellotti B, Rochecouste JF (2014) The development of conservation agriculture in Australia-
farmers as innovators. Int Soil Water Conserv Res 2:21–34. https://doi.org/10.1016/S2095-
6339(15)30011-3
14. Lunt ID, Eldridge DJ, Morgan JW, Witt GB (2007) A framework to predict the effects of
livestock grazing and grazing exclusion on conservation values in natural ecosystems in
Australia. Aust J Bot 55:401–415. https://doi.org/10.1071/BT06178
15. Eldridge DJ, Greene RSB, Dean C (2011) Climate change in the rangelands: implications for
soil health and management. In: Singh B, Cowie AL, Yin Chan K (eds) Chapter 13. Soil health
and climate change. Soil biology series. Springer, London, pp 237–255
16. Wasson RJ, Mazari RK, Starr B, Clifton G (1998) The recent history of erosion and sedimen-
tation on the Southern Tablelands of southeastern Australia: sediment flux dominated by
channel incision. Geomorphology 24:291–308. https://doi.org/10.1016/S0169-555X(98)
00019-1
17. Eldridge DJ, Poore AG, Ruiz-Colmenero M, Letnic M, Soliveres S (2016) Ecosystem structure,
function, and composition in rangelands are negatively affected by livestock grazing. Ecol Appl
26:1273–1283. https://doi.org/10.1890/15-1234
18. Eldridge DJ, Reed S, Travers SK, Bowker MA, Maestre FT, Ding J, Havrilla C, Rodriguez-
Caballero E, Barger N, Weber B, Antoninka A, Belnap J, Chaudhary B, Faist A, Ferrenberg S,
Huber-Sannwald E, Malam Issa O, Zhao Y (2020) The pervasive and multifaceted influence of
biocrusts on water in the world’s drylands. Glob Chang Biol 26:6003–6014. https://doi.org/10.
1111/gcb.15232
Agricultural Soil Degradation in Australia 65

19. Bullard JE, Ockelford A, Strong C, Aubault H (2018) Effects of cyanobacterial soil crusts on
surface roughness and splash erosion. Eur J Vasc Endovasc Surg 123:3697–3712. https://doi.
org/10.1029/2018JG004726
20. Williams W, Büdel B, Williams S (2018) Wet season cyanobacterial N enrichment highly
correlated with species richness and Nostoc in the northern Australian savannah.
Biogeosciences 15:2149–2159. https://doi.org/10.5194/bg-15-2149-2018
21. Eldridge DJ, Koen TB (2008) Formation of nutrient-poor soil patches in a semi-arid woodland
by the European rabbit (Oryctolagus cuniculus L.). Austral Ecol 33:88–98. https://doi.org/10.
1111/j.1442-9993.2007.01793.x
22. Hacker RB, Jessop PJ, Smith WJ, Melville GJ (2010) A ground cover-based incentive approach
to enhancing resilience in rangelands viewed as complex adaptive systems. Rangel J 32:283–
291. https://doi.org/10.1071/RJ10011
23. Braden J, Mills CH, Cornwell WK, Waudby HP, Letnic M (2021) Impacts of grazing by
kangaroos and rabbits on vegetation and soils in a semi-arid conservation reserve. J Arid
Environ 190:104526. https://doi.org/10.1016/j.jaridenv.2021.104526
24. Eldridge DJ, Ding J, Travers SK (2021) Low-intensity kangaroo grazing has largely benign
effects on soil health. Ecol Manage Restor 22. https://doi.org/10.1111/emr.12439
25. Robertson G, Wright J, Brown D, Yuen K, Tongway D (2019) An assessment of feral horse
impacts on treeless drainage lines in the Australian Alps. Ecol Manage Restor 20:21–30. https://
doi.org/10.1111/emr.12359
26. Chappell A, Sanderman J, Thomas M, Read A, Leslie C (2012) The dynamics of soil
redistribution and the implications for soil organic carbon accounting in agricultural south-
eastern Australia. Glob Chang Biol 18:2081–2088. https://doi.org/10.1111/j.1365-2486.2012.
02682.x
27. Orgill SE, Condon JR, Conyers MK, Morris SG, Alcock DJ, Murphy BW, Greene RSB (2018)
Removing grazing pressure from a native pasture decreases soil organic carbon in southern New
South Wales, Australia. Land Degrad Dev 29:274–283. https://doi.org/10.1002/ldr.2560
28. McDonald SE, Reid N, Smith R, Waters CM, Hunter J, Rader R (2019) Rotational grazing
management achieves similar plant diversity outcomes to areas managed for conservation in a
semi-arid rangeland. Rangel J 41:135–145. https://doi.org/10.1071/RJ18090
29. Byrnes RC, Eastburn DJ, Tate KW, Roche LM (2018) A global meta-analysis of grazing
impacts on soil health indicators. J Environ Qual 47:758–765. https://doi.org/10.2134/
jeq2017.08.0313
30. Hacker RB, Sinclair K, Pahl L (2020) Prospects for ecologically and socially sustainable
management of total grazing pressure in the southern rangelands of Australia. Rangel J 41:
581–586. https://doi.org/10.1071/RJ20006
31. Fleming PA, Anderson H, Prendergast AS, Bretz MR, Valentine LE, Hardy GES (2014) Is the
loss of Australian digging mammals contributing to a deterioration in ecosystem function?
Mamm Rev 44:94–108. https://doi.org/10.1111/mam.12014
32. Hamza M, Anderson W (2005) Soil compaction in cropping systems: a review of the nature,
causes and possible solutions. Soil Tillage Res 82:121–145. https://doi.org/10.1016/j.still.2004.
08.009
33. Walsh P (2002) New method yields a worm’s eye view. Farming Ahead 132:16–18
34. Roberton S, Lobsey C, Bennett JM (2021) A Bayesian approach toward the use of qualitative
information to inform on-farm decision making: the example of soil compaction. Geoderma
382:114705. https://doi.org/10.1016/j.geoderma.2020.114705
35. Bluett C, Tullberg JN, McPhee JE, Antille DL (2019) Soil and tillage research: why still focus
on soil compaction? Soil Tillage Res:1042382. https://doi.org/10.1016/j.still.2019.05.028
36. Zhang GS, Chan KY, Oates A, Heenan DP, Huang GB (2007) Relationship between soil
structure and runoff/soil loss after 24 years of conservation tillage. Soil Tillage Res 92:122–128.
https://doi.org/10.1016/j.still.2006.01.006
66 F. A. Dadzie et al.

37. Liu Q, Liu B, Zhang Y, Lin Z, Zhu T, Sun R, Wang X, Ma J, Bei Q, Liu G, Lin X, Xie Z (2017)
Can biochar alleviate soil compaction stress on wheat growth and mitigate soil N2O emissions?
Soil Biol Biochem 104:8–17. https://doi.org/10.1016/j.soilbio.2016.10.006
38. Phogat V, Cox JW, Šimůnek J (2018) Identifying the future water and salinity risks to irrigated
viticulture in the Murray-Darling Basin, South Australia. Agric Water Manag 201:107–117.
https://doi.org/10.1016/j.agwat.2018.01.025
39. Muyen Z, Moore GA, Wrigley RJ (2011) Soil salinity and sodicity effects of wastewater
irrigation in south-east Australia. Agric Water Manag 99:33–41. https://doi.org/10.1016/j.
agwat.2011.07.021
40. Hajkowicz S, Young M (2005) Costing yield loss from acidity, sodicity and dryland salinity to
Australian agriculture. Land Degrad Devel 16:417–433. https://doi.org/10.1002/ldr.670
41. Rengasamy P (2006) World salinization with emphasis on Australia. J Exp Bot 57:1017–1023.
https://doi.org/10.1093/jxb/erj108
42. ABS (2002) Salinity on Australian Farms, ABS Catalogue No. 4615.0 ISBN 0 642 47890 2.
Australian Bureau of Statistics, Canberra
43. Wong VNL, Dalal RC, Greene RSB (2008) Salinity and sodicity effects on respiration and
microbial biomass of soil. Biol Fertil Soils 44:943–953. https://doi.org/10.1007/s00374-008-
0279-1
44. Bañón S, Ochoa J, Bañón D, Ortuño MF, Sánchez-Blanco MJ (2019) Controlling salt flushing
using a salinity index obtained by soil dielectric sensors improves the physiological status and
quality of potted hydrangea plant. Sci Hortic 247:335–343. https://doi.org/10.1016/j.scienta.
2018.12.026
45. Boland A-M, Bewsell D, Kaine G (2006) Adoption of sustainable irrigation management
practices by stone and pome fruit growers in the Goulburn/Murray Valleys, Australia. Irrig
Sci 24:137–145. https://doi.org/10.1007/s00271-005-0017-5
46. Dang YP, Dalal RC, Buck SR, Harms B, Kelly R, Hochman Z, Schwenke GD, Biggs AJW,
Ferguson NJ, Norrish S, Routley R, McDonald M, Hall C, Singh DK, Daniells IG,
Farquharson R, Manning W, Speirs S, Grewal HS, Cornish P, Bodapati N, Orange D (2010)
Diagnosis, extent, impacts, and management of subsoil constraints in the northern grains
cropping region of Australia. Aust J Soil Res 48:105–119. https://doi.org/10.1071/SR09074
47. Page KL et al (2018) Management of the major chemical soil constraints affecting yields in the
grain growing region of Queensland and New South Wales, Australia – a review. Soil Res 56:
765–779. https://doi.org/10.1071/SR18233
48. Rocha F, Esteban Lucas-Borja M, Pereira P, Muñoz-Rojas M (2020) Cyanobacteria as a nature-
based biotechnological tool for restoring salt-affected soils. Agron 10:1321. https://doi.org/10.
3390/agronomy10091321
49. Feikema P, Baker T (2011) Effect of soil salinity on growth of irrigated plantation Eucalyptus in
south-eastern Australia. Agric Water Manag 98:1180–1188. https://doi.org/10.1016/j.agwat.
2011.03.005
50. Tian D, Niu S (2015) A global analysis of soil acidification caused by nitrogen addition.
Environ Res Lett 10:024019. https://doi.org/10.1088/1748-9326/10/2/024019
51. Hackney BF, Jenkins J, Powells J, Edwards CE, De Meyer S, Howieson JG, Yates RJ, Orgill SE
(2019) Soil acidity and nutrient deficiency cause poor legume nodulation in the permanent
pasture and mixed farming zones of south-eastern Australia. Crop Pasture Sci 70:1128–1140.
https://doi.org/10.1071/CP19039
52. Li GD, Conyers MK, Helyar KR, Lisle CJ, Poile GJ, Cullis BR (2019) Long-term surface
application of lime ameliorates subsurface soil acidity in the mixed farming zone of south-
eastern Australia. Geoderma 338:236–246. https://doi.org/10.1016/j.geoderma.2018.12.003
53. Condon J, Burns H, Li G (2021) The extent, significance and amelioration of subsurface acidity
in southern New South Wales, Australia. Soil Res 59:1–11. https://doi.org/10.1071/SR20079
54. Joffre R, Rambal S (1993) How tree cover influences the water balance of Mediterranean
rangelands. Ecology 74:570–582. https://doi.org/10.2307/1939317
Agricultural Soil Degradation in Australia 67

55. Caradus JR, Crush JR, Ouyang L, Fraser W (2001) Evaluation of aluminium-tolerant white
clover (Trifolium repens) selections on East Otago upland soils. NZ J Agric Res 44:141–150.
https://doi.org/10.1080/00288233.2001.9513470
56. Thai PK, Heffernan AL, Toms LL, Li Z, Calafat AM, Hobson P, Broomhall S, Mueller JF
(2016) Monitoring exposure to polycyclic aromatic hydrocarbons in an Australian population
using pooled urine samples. Environ Int 88:30–35. https://doi.org/10.1016/j.envint.2015.
11.019
57. McLaughlin MJ, Tiller K, Naidu R, Stevens D (1996) The behaviour and environmental impact
of contaminants in fertilizers. Soil Res 34:1–54. https://doi.org/10.1071/SR9960001
58. Loganathan P, Hedley M, Grace N, Lee J, Cronin S, Bolan N, Zanders J (2003) Fertiliser
contaminants in New Zealand grazed pasture with special reference to cadmium and fluorine – a
review. Soil Res 41:501–532. https://doi.org/10.1071/SR02126
59. Schöpfer L, Menzel R, Schnepf U, Ruess L, Marhan S, Brümmer F, Pagel H, Kandeler E (2020)
Microplastics effects on reproduction and body length of the soil-dwelling nematode
Caenorhabditis elegans. Front Environ Sci 8:41 10.18452/22474
60. Fergusson L (2017) Anthrosols and Technosols: the anthropogenic signature of contaminated
soils and sediments in Australia. Water Air Soil Pollut 228:269. https://doi.org/10.1007/s11270-
017-3460-z
61. Bradney L, Wijesekara H, Palansooriya KN, Obadamudalige N, Bolan NS, Ok YS, Rinklebe J,
Kim KH, Kirkham MB (2019) Particulate plastics as a vector for toxic trace-element uptake by
aquatic and terrestrial organisms and human health risk. Environ Int 131:104937. https://doi.
org/10.1016/j.envint.2019.104937
62. Mohajerani A, Karabatak B (2020) Microplastics and pollutants in biosolids have contaminated
agricultural soils: An analytical study and a proposal to cease the use of biosolids in farmlands
and utilise them in sustainable bricks. Waste Manag 107:252–265. https://doi.org/10.1016/j.
wasman.2020.04.021
63. AWA (2019) Australian biosolids statistics, Australian Water Association, Australia. https://
www.biosolids.com.au/guidelines/australian-biosolids-statistics
64. Ng E-L, Lwanga EH, Eldridge SM, Johnston P, Hu H-W, Geissen V, Chen D (2018) An
overview of microplastic and nanoplastic pollution in agroecosystems. Sci Total Environ 627:
1377–1388. https://doi.org/10.1016/j.scitotenv.2018.01.341
65. da Costa JP, Santos PS, Duarte AC, Rocha-Santos T (2016) (Nano)plastics in the environment-
sources, fates and effects. Sci Total Environ 566:15–26. https://doi.org/10.1016/j.scitotenv.
2016.05.041
66. de Souza Machado AA, Lau CW, Kloas W, Bergmann J, Bachelier JB, Faltin E, Becker R,
Görlich AS, Rillig MC (2019) Microplastics can change soil properties and affect plant
performance. Environ Sci Technol 53:6044–6052. https://doi.org/10.1021/acs.est.9b01339
67. Qi Y et al (2018) Macro-and micro-plastics in soil-plant system: effects of plastic mulch film
residues on wheat (Triticum aestivum) growth. Sci Total Environ 645:1048–1056. https://doi.
org/10.1016/j.scitotenv.2018.07.229
68. Ren Z, Gui X, Xu X, Zhao L, Qiu H, Cao X (2021) Microplastics in the soil-groundwater
environment: aging, migration, and co-transport of contaminants – a critical review. J Hazard
Mater 419:126455. https://doi.org/10.1016/j.jhazmat.2021.126455
69. Sobhani Z, Fang C, Naidu R, Megharaj M (2021) Microplastics as a vector of toxic chemicals
in soil: Enhanced uptake of perfluorooctane sulfonate and perfluorooctanoic acid by earth-
worms through sorption and reproductive toxicity. Environ Technol Innov 22:101476. https://
doi.org/10.1016/j.eti.2021.101476
70. Hughes L (2011) Climate change and Australia: key vulnerable regions. Reg Environ Change
11:189–195. https://doi.org/10.1007/s10113-010-0158-9
71. Arneth A et al (2019) IPCC special report on climate change and land. Summary for policy
makers. International Panel on Climate Change, Geneva, Switzerland
72. Linehan V, Thorpe S, Andrews N, Kim Y, Beaini F (2012) Food demand to 2050: opportunities
for Australian agriculture. ABARES conference, Canberra
68 F. A. Dadzie et al.

73. Hughes L, Steffen W, Rice M, Pearce A (2015) Feeding a hungry nation: climate change, food
and farming in Australia. Climate Council of Australia
74. Hughes N, Galeano D, Hattfield-Dodds S (2019) The effects of drought and climate variability
on Australian farms, Australian Bureau of Agricultural and Resource Economics and Sciences,
Canberra. CC BY 4.0. https://doi.org/10.25814/5de84714f6e08
75. Eldridge DJ, Beecham G (2018) The impact of climate variability on land use and livelihoods in
Australia’s rangelands. In: Squires VR (ed) Climate variability, land-use and impact on liveli-
hoods in the arid lands. Springer, New York
76. Rosenzweig C, Yang XB, Anderson P, Epstein P, Vicarelli M (2005) Agriculture: climate
change, crop pests and diseases. In: Epstein P, Mills E (eds) Climate change futures: health,
ecological and economic dimensions. The Center for Health and the Global Environment at
Harvard Medical School, pp 70–77
Agricultural Land Degradation in Peru
and Bolivia

Ronald R. Gutierrez, Frank Escusa, Miluska A. Rosas, and Mario Guevara

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
2 Levels of Soil Erosion in Agricultural Areas in Peru and Bolivia . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.1 The Amount of Soil Degradation in Agricultural Lands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.2 The Distribution of Soil Erosion in Agricultural Lands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3 Causes and Impacts of Soil Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.1 Agriculture Production Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.2 Soil Erosion and Salinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.3 Soil Erosion and Soil Organic Carbon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.4 Irrigated Agriculture Effect on Slope Stability in Peru . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.5 Pesticide Pollution Risk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.6 Soil Degradation and Microplastics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.7 Overgrazing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.8 Slash-and-Burn Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4 Solutions to Soil Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.1 The Role of Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.2 New Agriculture Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.3 Reforestation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.4 Erosion Control Regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

R. R. Gutierrez (✉) and F. Escusa


GERDIS, GEOSED, Departamento de Ingeniería, Pontificia Universidad Católica del Perú,
Lima, Perú
e-mail: [email protected]
M. A. Rosas
GERDIS, GEOSED, Departamento de Ingeniería, Pontificia Universidad Católica del Perú,
Lima, Perú
Université catholique de Louvainen, Ottignies-Louvain-la-Neuve, Belgium
M. Guevara
Centro de Geociencias, Universidad Nacional Autónoma de México, Campus Juriquilla,
Querétaro, México

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 69
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 69–96, DOI 10.1007/698_2022_926,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 14 December 2022
70 R. R. Gutierrez et al.

Abstract Soil erosion is the most prominent evidence of soil degradation and is
regarded as the main type of agricultural land degradation worldwide. The objective
of this contribution is to assess soil degradation associated with agricultural lands in
Peru and Bolivia. The assessment is mainly based on a recent soil erosion survey;
maps describing soil erosion rates across 5 km grids for the years 1990, 2000, 2010,
and 2020; and global observations of agricultural land areas. We observe that soil
erosion rates in agricultural lands from Peru and Bolivia are steadily growing. We
also observe that (1) the highest quantities of soil erosion are found in the Andean
highlands and the largest areas of soil denudation in the Amazonian sections, and
(2) there is a positive relation between erosion degree and soil organic carbon stock.
The main drivers of soil erosion are the intensive agriculture production and fertilizer
use that have increased in Peru and Bolivia. This land use intensification in both
countries is also the driver of other land degradation factors such as soil salinization
and soil pollution. The use of sewage (e.g., gray water) in urban and peri-urban
agricultural areas increases the risk to soil health by inducing the dispersion of
microplastics assisted by soil erosion. Overgrazing affects a large portion of the
highlands from both countries. Very limited information of the extent of slash-and-
burn fire agriculture exists; however, this practice is mostly observed in the Depart-
ment of Santa Cruz (Bolivia). To prevent or minimize the pressure on soil resources
associated with soil erosion or soil pollution, we highlight the need to support
sustainable soil management practices and responsible use of fertilizers in
agricultural land.

Keywords Agriculture, Bolivia, Fire, Overgrazing, Peru, Slash and burn, Soil
degradation

1 Introduction

Soil degradation is controlled by physical (e.g., deterioration of soil structure,


compaction, erosion, desertification), chemical (e.g., depletion of soil fertility, acid-
ification, salinization, and pollution), and biological processes [1]. These processes
are driven by natural causes (e.g., steep slopes, floods, storms, droughts) and
anthropogenic factors including deforestation, overexploitation of vegetation,
overgrazing, excessive use of agrochemicals, and poor conservation practices [1].
The most significant causes of soil degradation worldwide are erosion, decline of
soil organic carbon (SOC) and biodiversity, nutrient imbalance, salinization, con-
tamination, compaction, surface sealing, and waterlogging [2, 3]. The degree of
severity, spatial extent, and the interaction among these causes are nevertheless
complex and diverse [2]. In the last few years, there is also a growing concern
about the presence of macro- and microplastics in soils and the effect on soil-plant
system [4].
Agricultural Land Degradation in Peru and Bolivia 71

From the causes in question, soil erosion is regarded as the main type of soil
degradation worldwide [5]. Thus, there is global societal concern over soil erosion,
particularly in tropical developing nations, where it is reaching critical rates [6, 7]. In
Latin America, for example, nearly 20% of land area is being affected by soil erosion
[8, 9]. This is due to anthropogenic stressors such as a marked increase of land used
for agriculture [10] and the impacts of agronomic technical change, inappropriate
government policies, and poor institutional quality [11].
Erosion directly affects on-site agronomic productivity and rises off-site costs
associated with the operativity of hydraulic infrastructure (e.g., reservoirs, water
ways, flood protection works) due to siltation. Water treatment also increases due to
sediment concentration increase [12, 13]. Bolivia (a lower-middle income economy)
and Peru (an upper-middle economy) have increased their agricultural production in
the last two decades. Although Bolivia exhibits a high proportion of small-scale
farmers, in the last two decades, flex crops – namely, crops that have multiple uses
that can be easily and flexibly inter-changed – have expanded the Bolivian agricul-
tural production mainly through deforestation [14–16]. In 1993, the Peruvian gov-
ernment issued explicit policies to stimulate investment and the adoption of new
technologies in the agricultural sector. It allowed the country for exhibiting the
greatest annual production growth (5.6%) in the Latin American region between
1993 and 2008, and to maintain a growing trend in the last two decades [17, 18].
Current estimates indicate that the Andean catchments from Peru and Bolivia
combined are the largest contributors of sediments for the Amazon system
[19]. Higher sediment rates from Bolivian catchments than those from Peru were
observed in previous studies [20]. Likewise, due to a potential expansion of cropland
areas in Bolivia and Peru, soil erosion, and consequently discharge of particulate
organic carbon, may increase in the upcoming years [6, 21]. In addition, these
countries are included among the top 30 nations exhibiting high risks for soil
degradation by fertilizers contamination [22]. To date, accessible studies on soil
degradation at country scale for Peru and Bolivia are limited or not available. Most
studies report area-specific analysis of soil degradation in severely affected areas
[7, 23, 24]. Moreover, based on our academic and research experience in Peru and
Bolivia, we believe that soil erosion appears far less in public discourse compared to
other topics such as water pollution, which is also a pattern commonly observed in
many countries [10].
Rosas and Gutierrez [7] highlighted the technical challenges of quantifying soil
erosion rates in developing countries, which commonly exhibit limitations to build
and validate soil erosion models. They also introduced RUSLE-GGS, a GIS-based
RUSLE methodology, to quantify soil erosion rates at country scale applicable to
these nations. RUSLE-GGS allows for obtaining soil erosion rates estimates with a
known uncertainty and has been successfully applied to Peru.
The main objectives of this study are (1) conducting a digital soil degradation
assessment (e.g., salinization, SOC loss, soil erosion) based on freely accessible
72 R. R. Gutierrez et al.

global data (e.g., global food security-support analysis data,1 global distribution of
field size2), past studies, and soil erosion rates estimates obtained through the
RUSLE-GGS methodology by Rosas and Gutierrez [7]; (2) reviewing existing
evidence of the severity and concentration of slash-and-burn agriculture and
overgrazing; and (3) identifying potential actions to improve the knowledge and
the effects of soil degradation. This study could be used by decision-makers, public
agencies, and scientists from Peru and Bolivia as a knowledge resource to improve
the understanding of the dynamics of soil degradation through a multidisciplinary
approach.

2 Levels of Soil Erosion in Agricultural Areas in Peru


and Bolivia

This section examines the current condition of soil erosion in Peru and Bolivia
because it is regarded as the main type of soil degradation worldwide and a major
risk to soil health in Latin America [8]. That being so, it is analyzed by applying the
RUSLE-GGS methodology by Rosas and Gutierrez [7]. RUSLE-GGS comprises the
following steps to estimate soil erosion rates: (1) construction of erosion rate samples
using RUSLE-GIS based on freely available global geoenvironmental observations
(e.g., NASADEM Merged DEM Global 1 arc sec [25], NOAA NDVI version 4 [26],
SoilGrids Global gridded soil information [27], CHIRPS Daily Precipitation with
Station Data [28]), model outputs, and field relations; (2) construction of area-
specific sediment yield samples utilizing a set of soil delivery ratio functions; and
(3) assessment of the most behavioral samples through bias analysis and cross
validation.
The application of RUSLE-GGS to Peru [7] for 1990, 2000, and 2010 indicates
that the most behavioral erosion rate samples are predominantly obtained by means
of the following fundamental parameters of RUSLE: the Arnoldus rainfall erosivity
factor based on satellite precipitation input data, cover and management factor based
on global data, and slope length/steepness factor based on the LS-TOOL output [29],
and the sediment delivery ratio function by Williams and Berndt [30]. The erosion
rate estimates circumscribed to agricultural lands are obtained by using the global
observations thereof for the year 2000 [31]. The quantification of erosion rates in
agricultural lands provides the basis for the subsequent sections, in which the causes,
consequences, and solutions to soil degradation are discussed.

1
https://croplands.org/app/map?lat=0.17578&lng=0&zoom=2 (Accessed on 03/10/2021).
2
https://pure.iiasa.ac.at/id/eprint/15526/ (Accessed on 06/05/2021).
Agricultural Land Degradation in Peru and Bolivia 73

Table 1 Erosion coverage evolution by degree in agricultural lands from Peru and Boliviaa
Peru Bolivia
1990 2000 2010 2020 1990 2000 2010 2020
(%) (%) (%) (%) (%) (%) (%) (%)
Very low 17.0 12.9 10.5 10.0 67.3 55.1 45.7 60.1
Low 38.0 31.5 25.1 25.1 20.7 31.8 39.1 25.4
Moderate 20.8 21.3 19.4 19.2 6.2 6.1 5.9 5.5
low
Moderate 13.1 14.8 16.0 16.3 3.1 3.6 4.2 4.1
High 6.9 10.9 14.2 13.6 2.1 2.3 2.9 2.7
Severe 4.2 8.5 14.8 15.7 0.6 1.1 2.2 2.3
a
These quantities should be regarded as a broad erosion rate survey for Peru and Bolivia

2.1 The Amount of Soil Degradation in Agricultural Lands

Peru and Bolivia exhibit similar biomes in their Amazonian and Andean sections
[32–34]; pollution characteristics (e.g., the largest percentages of population using
solid fuels along the Andes), long-term climatology and future potential impacts
associated with global warming, and increase in detected fires [35]; and economic
reliance on extractive industries such as mining [36]. Therefore, it is reasonable to
assume that the aforementioned RUSLE structure to estimate erosion rates for Peru
can be extended to Bolivia. Thus, erosion rates estimates for Bolivia for 1990, 2000,
2010, and 2020, as well as those for Peru for the year 2020, are obtained. These
results should be regarded as a broad-scale erosion survey for assessing erosion rates
spatio-temporal evolution [7].
Table 1 summarizes our results, which are categorized according to a 6-class
scheme of soil erosion degrees, namely: very low (<1 ton/ha/year), low (<10 ton/ha/
year), moderate low (<25 ton/ha/year), moderate (<50 ton/ha/year), high
(<100 ton/ha/year), and severe as any rate above 100 ton/ha/year [3]. As shown in
Table 1, there is an increasing overall trend of erosion rates in Peru and Bolivia, with
Peru being the most affected by presenting in 2020 an increased agriculture area with
severe soil degradation of about 3.7 times the modeled coverage in 1990. Bolivia
exhibits a similar growth in degradation, although its base values are about 15% of
Peru.

2.2 The Distribution of Soil Erosion in Agricultural Lands

The distribution of agricultural land based on the Köppen climate units by Peel et al.
[37] shown in Fig. 1 indicates that the largest area of agricultural land in Peru and
Bolivia develops in the tundra climate (ET) and tropical savannah climate (Aw),
respectively.
74 R. R. Gutierrez et al.

Fig. 1 Land coverage distribution of agriculture areas from Peru (left) and Bolivia (right) based on
[31] and according to the Köppen climate classification [37]. Af tropical rainforest climate, Am
tropical monsoon climate, Aw tropical savannah climate, BSh hot semi-arid climate, BSk cold
semi-arid climate, BWh hot desert climate, BWk cold desert climate, Cfb temperate oceanic
climate, Csb warm Mediterranean climate, Cwa monsoon humid subtropical climate, Cwb subtrop-
ical highland climate, Cwc cold subtropical highland climate, ET tundra climate

Overall, Fig. 2 indicates that severe erosion in agricultural lands occupies a larger
proportion of the Peruvian territory than in the Bolivian one. This suggests that
tropical rainforest, cold and hot semi-arid, and tundra climate units in Peru and
Bolivia are persistently affected by severe erosion in 2020. The highest proportion of
severe erosion in Bolivia was identified in the subtropical highland climate (~15%).
In Peru most of severe erosion (~55%) was observed in agricultural lands from the
highlands.

3 Causes and Impacts of Soil Degradation

This section surveys the potential causes of soil degradation by agricultural produc-
tion in Peru and Bolivia. To that end, the recent temporal dynamics of agricultural
production, its associated water depletion potential, and fertilizer consumption in
these countries are analyzed. Then, a digital assessment of the relationship of these
dynamics with the spatio-temporal evolution of erosion rate, salinity, and organic
carbon is performed. This assessment also focuses on the pesticide pollution risk, the
increasing presence of microplastics on soils, and, for the specific case of Peru, the
potential causes of slope instability associated with irrigated agriculture. Finally, the
potential effect of overgrazing slash-and-burn agriculture in both countries is
discussed.
Agricultural Land Degradation in Peru and Bolivia 75

Fig. 2 Severe erosion distribution in agricultural lands per climate unit based on global agricultural
land maps for the year 2000 by Ramankutty et al. [31]. Figure 1 describes the acronyms

3.1 Agriculture Production Dynamics

According to FAO [38], over the last two decades, the crop production and the use of
fertilizers such as nitrogen (N), phosphate (P2O5), and potash (K2O) increased over
the last few decades in Peru and Bolivia (Fig. 3). The most evident increase was
observed in maize and potato production, which are the most representative crops in
these countries. As observed in Fig. 3, the Peruvian production increased rapidly
from the year 1990 onward due to constitutional changes introduced in the early
1990s that removed restrictions to private ownership of land, liberalized land
markets, and incentivized private acquisition of land in the Coastal region [17]. Sim-
ilarly, Bolivia also exhibits a growing production and fertilizer use after the 2000s
76 R. R. Gutierrez et al.

Fig. 3 Representative crop production of Peru and Bolivia compared to the usage of fertilizers in
tons [38]

(Fig. 3). It possibly responded to a growing foodstuff internal demand as the country
has historically prioritized an agricultural autarky system [39].
Hofste et al. [40] estimated the worldwide water depletion and stress-inducing
hazard by comparing all potential water sources and the water demand from different
uses. In Peru and Bolivia, agricultural production is the largest consumer of water. In
these countries, the main factor contributing to water scarcity is the overexploitation
of water for agricultural production, mainly across water-limited environments in the
arid coastal region (where large-scale irrigated agriculture develops) and the low
Andean region. Figure 4 represents the agriculture land distribution and the areas
with the water depletion potential.

3.2 Soil Erosion and Salinity

Soil salinity affects crop production as it reduces plant growth and induces crop
wilting by interfering with N uptake. Thus, several techniques (e.g., use of transgenic
breeds or bacteria-enriched fertilizers) were proposed to fight high salinization
[41]. However, these methods only inhibit the immediate effects on current crops
without reducing the potential of long-term salinity transport [42].
A global satellite-based study by Ivushkin et al. [43] for the period 1986–2016
identified an important saline concentration in the Piura region soil (Northern
Peruvian Coast in Fig. 5), which is confirmed by local studies that indicate the
increase of salinity due to unregulated rice crops that are mainly irrigated by flooding
the terrain [44]. Ivushkin et al. [43] did not identify any important soil salinity
hotspot for Bolivia, possibly due to the limitations of the methodology. Neverthe-
less, preliminary results from regional soil salinity assessments led by the Global
Partnership of the Food and Agriculture Organization [45] suggest that other impor-
tant areas with high soil salinity, mainly in Arequipa in Peru and La Paz or in the
higher land surrounding the Titicaca Lake in Bolivia. These soil salinity patterns
Agricultural Land Degradation in Peru and Bolivia 77

Fig. 4 Overall water depletion potential in Peru and Bolivia in contrast to the water stress induced
exclusively by agricultural practices based on freely accessible supporting material provided by
Hofste et al. [40]

were associated with the increase of highly irrigated areas shown in Fig. 5 and
reported by Lacroix et al. [46].
Figure 5 combines erosion degree, categorical information of salinity hotspots,
and average regional (i.e., geographical unit called region in Peru and department in
Bolivia) SOC stocks [27]. A visual inspection suggests that there is a positive
relation between erosion degree and salinity in the Peruvian South West section. A
high soil salinity is also observed in the arid Northern coastal section, which may
also increase in the coming years if the agricultural expansion continues. In the
Bolivian territory, salinity appears to be reaching slightly saline degrees. The only
extreme saline spot is in the vicinity of Uyuni salt flat.
78 R. R. Gutierrez et al.

Fig. 5 Comparison of the erosion degree in 2010 with the soil salinity at surface level and organic
carbon stock down to 30 cm depth per region based on freely available data provided by Hengl et al.
[27]. Salinity categories are defined as follows: non-saline (<2 dS/m), slightly saline (<4 dS/m),
moderately saline (8 dS/m), highly saline (16 dS/m), and extremely saline (>16 dS/m). Thin gray
lines represent the limits of geographical units called regions in Peru and departments in Bolivia

3.3 Soil Erosion and Soil Organic Carbon

A key component of a healthy and fertile soil is SOC, which is the major component
of organic matter. SOC serves as a key nutritional resource for plants and microbes.
Although Earth’s soils are a rich reservoir of carbon, both soil erosion and current
agricultural practices have a predominantly detrimental impact on SOC, especially
on sloping Andean areas where the SOC loss due to erosion is more pronounced.
Agricultural practices such as intensive tillage, used in these areas, cause declines of
SOC stocks due to oxidation of organic matter, destruction of soil aggregates, and
reduction in water infiltration [47].
Agricultural Land Degradation in Peru and Bolivia 79

At the regional scale, the interaction between soil erosion degree and SOC stock
is observed in Peru and Bolivia (Fig. 5). The northern part of Peru is identified as a
tropical alpine environment, which represents an important SOC reservoir. The
continuous vegetation cover, low air temperature, low atmospheric pressure, and
frequent soil waterlogging increase SOC accumulation [48, 49]. However, many
tropical alpine areas present considerable human activity. Due to high pressure for
arable land, land use changes are observed from alpine grasslands to croplands rather
than forest [50]. SOC storage under these land uses is reduced due to the lack of
fertilization and manuring combined with high levels of erosion [51]. The tropical
and subtropical moist broadleaf forest in the Amazonian region of Peru and Bolivia
is experiencing high deforestation rates [52]. Along with deforestation, the area is
affected by soil erosion, which occurs when heavy rains fall on deforested bare soil
with gentle and moderate slopes. Soil erosion disrupts the continuous addition of
organic matter to the soil, altering SOC dynamics, affecting the soil biodiversity, and
consequently affecting soil ecosystem functions [53].
The arid region of the Peruvian coastal plain has the lowest SOC stocks of Latin
America and the Caribbean with rates under 25 ton C/ha (Fig. 6). A previous study
[54] estimated a rate of SOC loss of around 15 ton C/ha. Similarly, Bolivia exhibits
some regions with low soil carbon content (25–75 ton C/ha) in the Eastern Andes
and very low (<25 ton C/ha) in the Western Andes (Fig. 5). In these arid regions, it
becomes essential to preserve the scarce carbon due to the fragility of the local
ecosystems [55]. Continuous SOC losses across these areas may trigger irreversible
land desertification processes.

3.4 Irrigated Agriculture Effect on Slope Stability in Peru

Both rainfall-driven and internal soil erosion play an important role in slope stability
[56, 57]. The latter includes two phenomena: suffusion (i.e., fine particles are eroded
through the voids between larger particles by seepage flow, leaving behind a coarse
soil skeleton) and piping (i.e., progressive erosion and transportation of soil particles
along a flow path), which diminishes soil stability, mainly in cohesionless soils as
silt and silica sands [58].
Large-scale and irrigation-intensive agriculture led to increased volumes of
seepage in some coastal Peruvian areas [59]. For instance, in the Quilca watershed
located in the coast of the Arequipa region (see location in Fig. 5), the constant
manipulation and deterioration of the existing soil due to the farming practices
induced several landslides over the last 40 years [42, 46]. A similar scenario was
also observed in the mid/upper Jequetepeque watershed (Cajamarca region in
Fig. 5), where small-scale irrigation interventions were enacted to promote organic
methods for fruit production and pasture [60].
80 R. R. Gutierrez et al.

Fig. 6 Pesticide risk score, number of active ingredients of pesticides, and overall concern level of
pesticide contamination potential [22] coupled with the erosion rate estimates for 2020 displayed as
elevations. The circle and red arrow highlight the location of the high concern level in the
Jequetepeque area. Spikes represent local peaks of erosion rate

3.5 Pesticide Pollution Risk

Figure 6 shows the relationship between erosion rates for 2020 (displayed as
topographical information) and the pesticide risk score, number of pesticide active
ingredients (i.e., pesticide use), and concern level due to potential impact on the
existing biodiversity for the year 2021 by Tang et al. [22].
From Fig. 6, although not quantified herein, it is reasonable to infer that there is a
positive relation between erosion rate and pesticide risk score and erosion rate and
the number of active ingredients. This is particularly evident in Andean highlands of
Peru and Bolivia and some possibly deforested areas in the Amazonian section. Most
of the study area is nevertheless categorized below the medium concern level with
Agricultural Land Degradation in Peru and Bolivia 81

Fig. 7 Land fraction exposed to certain pollution risks. HWR = low and variable water supplies
exposed to high pesticide risk, HB = significant biodiversity exposed to pesticides directly,
EA = endangered amphibian species critically endangered by pesticides (based on freely accessible
material provided by Tang et al. [22])

only one exception, namely, the Jequetepeque region (Northern Peruvian Coast),
which presents a high level of concern. As underlined above, this region also
exhibits high salinity [44].
Tang et al. [22] ranks the World’s 30 countries exhibiting the highest levels of
pollution risk associated with pesticides based on three criteria: (1) water supplies
exposed to high pesticide risk, (2) significant biodiversity exposed to pesticides, and
(3) endangered amphibian species critically endangered by pesticides. Comparing
multiple South American countries (Fig. 7), Peru, along with Ecuador and Mexico, is
exposed to risks associated with the three criteria. Bolivia is affected by risks
associated with the first and the second criterion (Fig. 7).

3.6 Soil Degradation and Microplastics

Microplastics are important factors of soil degradation with negative implications for
soil biodiversity. Recent studies suggest that microplastics affect soil microbial
activity [61] and the microbiota of soil animals such as centipedes [62]. They also
inhibit plant growth in wheat crops [4]. These facts raise concerns on the increasing
presence of microplastics in soils across Bolivian and Peruvian territory over the last
few years. For instance, high-density microplastic traces are observed in coastal and
fluvial environments [63–65].
Certain microplastics affect the soil mechanical properties. For example, poly-
esters commonly found in clothing reduce the soil bulk density of loamy sands and
induce the formation of soil clumps [61]; other microplastics may affect overall
slope stability and erodibility rate in soils [57]. Thus, microplastics could exacerbate
soil erosion [61], which is one of the main vehicles for microplastic spread
82 R. R. Gutierrez et al.

[65]. Certainly, further research of this processes will be needed on erosion-prone


areas from the Andean sections of Peru and Bolivia.
Sewage is an important carrier of microplastics to the soil and freshwater
[66]. Therefore, there are calls to diminish the usage of sewage sludge as a fertilizer
alternative [67]. Sewage-derived waters are used in urban and peri-urban farms at
large Peruvian cities and several areas of Bolivia [68, 69], which may be locally
increasing the concentration of microplastics in soil.
Some initiatives for microplastics control exist already in Latin America. For
instance, the usage of bacteria to eliminate low-density polymers and the tests
performed in the Cancharani landfill (Puno, Peru) suggest promising results
[70]. Possibly, these initiatives must be taken with care, as some plastic eating
bacteria (e.g., Pseudomonas aeruginosa) is lethal to fauna and even to humans.

3.7 Overgrazing

Overgrazing mainly results from grass cover depletion by livestock. It inhibits


natural soil regeneration, promotes seed dispersal of ruderal species, and increases
soil compaction/trampling in areas where non-palatable leaves grow
[71]. Overgrazing is widespread in the Peruvian highlands (ET in Fig. 2), where
90% of the country’s cattle are raised [72]. In the absence of a conservation and
protection regulation of headwaters before 2017, overgrazing impacted the inherent
fragile ecosystem of some headwater areas [73–75]. However, there is some evi-
dence that the impact of overgrazing is to some extent attenuated by rotational
grazing practices conducted by poorer farmers [76, 77].
Overgrazing also severely affects some nonconverted tropical dry forest located
in the northwestern coastal region of Peru (Bwh in Fig. 2), which is mainly caused by
free-roaming cattle, goats, and donkeys [71, 78]. In this geographical context,
apparently overgrazing poses higher threats for some tree populations than climate
change [71].
Overgrazing impacts a large portion of the Bolivian Andes too. Although there is
a limited number of studies on the matter, they suggest that it is concentrated in the
highest areas and started to grow since the 1950s, possibly due to population, cattle
increase, and the lack of forage alternatives in dry years [79–83]. Recent research
reports that overgrazing is affecting highland wetlands or (bofedales), thus altering
headwaters hydrologic response [84–86].

3.8 Slash-and-Burn Agriculture

Slash-and-burn agriculture is mainly used by small-scale forest farmers to forage and


regenerate land productivity of previously cleared areas through several fallow
cycles [87]. Currently, there are limitations in the estimation of the extent of slash
Agricultural Land Degradation in Peru and Bolivia 83

and burn, its ecological impacts, and the rate at which it is changing in Amazonian
countries such as Peru and Bolivia [87].
In Peru, slash-and-burn increase is mainly triggered by economic structures such
as credit or cooperatives and public policies; likewise, slash-and-burn decrease is
controlled by economic structures and public attitudes toward it [87]. It is used
chiefly by indigenous farmers on both upland forest sections and riverine corridors,
mainly for cattle and agricultural production [17, 88, 89]. Anecdotical evidence
suggests that slash and burn does not necessarily exhibit a simple dynamic. For
instance, in the vicinity of Iquitos (the largest city in the Peruvian forest), it started to
increase since the 1950s, accelerated its pace in 1980s reaching a peak in the early
1990s, and then started to decrease [90]. Overall, it appears that slash and burn is one
of the main causes of forest and land degradation [89, 91, 92] and that it may
marginally affect the water chemistry of streams [93].
Slash and burn is practiced in most of the Bolivian forest section [94]. Neverthe-
less, most of the research on its use concentrates in the department of Santa Cruz,
which, at this point, occupies 61% of the national crop area [95] largely obtained
through slash-and-burn cultivation [96, 97]. In the late 1980s, slash and burn in
Santa Cruz was exacerbated by structural adjustment policies that resulted in a long-
term increase of the crop area and degradation of fragile soils [98, 99]. Slash-and-
burn agriculture is observed in both piedmont and alluvial areas from Santa Cruz
[100]. In some areas, soils exhibit the presence of charcoal and pottery shards [101].

4 Solutions to Soil Degradation

This section identifies some potential solutions to alleviate soil degradation. It


underscores the opportunities of adopting new technology and new agricultural
techniques, the need of a better understanding of the role of soil conditions and
trees restoration, and the urgency of establishing regulatory frameworks for erosion
control in Peru and Bolivia.

4.1 The Role of Technology

Precision agriculture emerged to optimize agricultural practices with the use of


several tools and devices. The most used in this field are remote sensors, followed
by UAV monitoring and mapping [102]. These two technologies are already in use
in Peru and Bolivia [103, 104]. In Peru, this is thanks to the effort of several public
agencies and private large-scale agriculture companies [105].
Agriculture sensors (e.g., proximal soil sensing) allow farmers to obtain real-time
observations of the soil response to certain parameters such as temperature, humid-
ity, and even greenhouse gas emission. Models and systems designs for soil sensing
sensors vary from large-scale monitoring and real-time server diagnostic tracking
84 R. R. Gutierrez et al.

[106] to small-scale farming where the sensors may only trigger immediate warning
to farmers via cellphone when a potential hazard is detected [107]. These practices
were driven by the Peruvian Ministry of Environment to improve the crop quality in
Lambayeque, Peru, yielding favorable results [104]. Likewise, prototypes tailored
for banana and mango crops in Sullana, Peru, demonstrated the effectiveness and
efficiency of these methods cost wise [108]. In Bolivia, up-to-date technology is
commonly used in export crops (e.g., soybean, sunflower, sugar) and fruit trees such
as bananas and grapes [39]. However, it is being adopted at a low pace due to the
shortage of well-trained personnel to operate the associated equipment and
devices [103].
Drones, on the other hand, are small aircrafts often controlled by a remote control
or a pre-programed routine which carry a payload, which is often a camera or
infrared sensor, to gather soil data and soil information from inaccessible areas
[109]. Although it is not widely reported, this method is gaining popularity in Peru
and Bolivia as it allows for easily assessing several parameters from large areas of
crops, such as plant health monitoring, canopy cover mapping, field water ponding,
and chlorophyll measurement, among others [110]. These examples of remote
monitoring tools are becoming essential to agricultural practices in Peru [105] and
Bolivia [103]. Possibly they will also be used for real-time erodibility sensing as the
capabilities of particle tracking sensors improve [111], which may advance the
practical assessment and the scientific understanding of soil degradation dynamics
in developing countries.

4.2 New Agriculture Techniques

Aside from the precision farming techniques discussed in Sect. 4.1, Agriculture 4.0,
a production system that will not depend on the uniform and massive use of water,
fertilizers, and pesticides across entire fields [112, 113], has been introduced in
recent years. Agriculture 4.0 will involve 3D printing of foods, cultured meat,
genetic modification, and seawater agriculture, which could grow in arid areas by
using abundant and clean resources such as the sun and seawater and new sophis-
ticated technological elements such as artificial intelligence, robots, temperature and
moisture sensors, aerial images, and GPS [112]. The arid Peruvian coastal region,
where most of the country population is concentrated, possibly has a high potential
for the application of large-scale Agriculture 4.0, especially in large cities (e.g.,
Lima, Trujillo, Piura) peri-urban areas.
Vertical farming or indoor farming represents another alternative to conventional
farming as it directly addresses the areal limitations thereof [114]. In contrast to
traditional farming, vertical farms require sophisticated equipment and a highly
specialized workforce [115]. Likewise, they are energy-intensive, especially those
using hydroponics that also require a relatively large number of personnel to
properly function [116]. Nevertheless, if planned correctly, vertical farming systems
can vastly outperform traditional farming, reporting returns of investments of 20%
Agricultural Land Degradation in Peru and Bolivia 85

and doubling the crop yield of leafy green vegetables under certain conditions
[115]. Most of the development in this field points to aeroponic cultivation as it
markedly reduces resources consumption [117].
The potential environmental benefits (e.g., reduction of fertilizers and water use)
of large-scale application of vertical agriculture have been identified in both devel-
oped and developing nations [118, 119]. By extension, in the geographical context
of Peru and Bolivia, it would be central to reduce the rate of deforestation and
expansion of agricultural frontiers. It would also reverse the increasing need to build
small- and large-scale irrigation reservoirs, create profitable small and medium-size
peri-urban agricultural business, decrease the pressure over soil, and improve bio-
diversity. However, these potential benefits could represent additional challenges to
smallholder farmers (the weakest group in the agricultural production chain), par-
ticularly those from remote high-altitude Andean and Amazonian regions of Peru
and Bolivia.

4.3 Reforestation

Reforestation is often seen as a viable nature-based solution to tackle soil erosion.


The plant root systems have the capacity of binding soil particles and increasing soil
porosity, which results in an important reduction of soil erosion vulnerability and
enhancement of water infiltration capability [120]. Despite these benefits, the eco-
logical knowledge of the role of soil conditions and altitude on trees restoration
potential remains scarce and patchy in central Andean and Amazonian regions [121–
125].
In general, forest and landscape restoration projects follow four principles:
enhancing and diversifying local livelihoods, not replacing native tropical grasslands
and savanna ecosystems, promoting landscape heterogeneity and biological diver-
sity, and distinguishing residual carbon stocks from new carbon stocks [126]. For
mountain regions with a steep slope and periodic storm events, as in the Peruvian
and Bolivian highlands, the most viable methods are a combination of mountain
terraces and the inclusion of deep-rooted trees as pines or alder. However, these
species have natural limitations to adapt to high slopes and require particular care as
their survival relies on the availability of fertile soil [127]. In contrast, in desert
regions, such as the Western Peruvian Andes, local varieties of shrubs may be
utilized to accumulate enough moisture to strengthen and control soil erosion by
the wind [128].
Reforestation practices must analyze the fact that inclusion of foreign species may
disrupt the local biodiversity and disturb the local carbon stocks [126]. Furthermore,
proposed solutions need to consider the prevalent erosion source because high water
retention may lead to an increment in pore pressure, which can potentially cause
large-scale landslides in regions where the erosion is rainfall driven rather than wind
driven [129].
86 R. R. Gutierrez et al.

Small-scale illegal mining has severely impacted large areas of fragile ecosystem
from the Peruvian and Bolivian Amazon at an accelerated pace [130, 131]. Appar-
ently protecting regulation has been more effective on the Peruvian side [131–
133]. Possibly it is nevertheless fair to state that a wide knowledge gap on the
technical approaches for reforestation of these areas still exists [121, 122, 134].
Brancalion and Chazdon [126] highlight the necessity of accounting financial
assessment and modeling prior to any reforestation initiative. Unfortunately, in many
cases in the Andean region, these projects are poorly executed or abandoned due to
budget constraints or lack of government incentives, thus undermining the original
purpose, and affecting the economy of the local communities [60].

4.4 Erosion Control Regulation

There is a need of an adequate soil governance frame (i.e., laws, policy, and law
enforcement) for soil erosion control in developing countries [7, 11, 135, 136]. Ero-
sion has not only largely contributed to soil degradation in Peru and Bolivia, but also
it seems that it has exacerbated poverty in these countries [80, 137]. Currently, Peru
and Bolivia lack specific regulation for controlling soil erosion; they only have
general guidelines and recommendations on the matter (Table 2), which are com-
monly strictly verified in Peruvian mining and petroleum operations [138].
We postulate that in designing and enacting soil erosion control regulations, the
countries in question could consider the requirements governing land-disturbing
activities from developing countries. For instance, under the US Environmental
Protection Agency’s National Pollutant Discharge Elimination System, a soil ero-
sion and sediment control plan should be developed for construction activities that
disturb one acre (0.4 ha) or more, or smaller projects that are part of a larger common
plan of development [140]. Likewise, according to the policies of the US Federal
Highway Administration, since 2016 “all highway projects should be located,
designed, constructed and operated according to standards that will minimize soil
erosion and sediment damage to the highway and adjacent properties and abate
pollution of surface and ground water resources” [141]. In addition, the US Depart-
ment of Agriculture establishes general soil erosion control treatment categories
(e.g., slope grading; seed, fertilizer, and soil amendments; soil stabilizers and
tackifiers; mulching) and promotes the use of nature-based solutions [142]. The
efficiency of some of these measures, however, has been questioned in recent
studies [143].

5 Conclusions

This contribution presents a novel digital assessment of soil degradation in Bolivia


and Peru. It is based on a large literature review, country-scale historical estimates of
soil erosion rates (1990, 2000, 2010, and 2020) obtained by applying RUSLE-GGS
Agricultural Land Degradation in Peru and Bolivia 87

Table 2 List of laws and regulations regarding erosion mitigation and control in Peru and Bolivia
Country Codea Name Article Description
Peru Ley N° Law that declares of national Art. 1 It is of national and public
30.557 interest and public necessity interest to improve river
the construction of flood defense structures, which will
defenses and hydraulic be included in the national
structures management agenda
Art. 2 The central government will
manage the project coordina-
tion and budget related to
flood defenses
Ley N° Law that establishes three Art. 3 Regarding agriculture, the
30.191 measurements for prevention, scope of this law will manage
reduction, and readiness in the the budget of river defenses,
presence of a disaster highland watershed treat-
ment, riverbank restoration
and control, and ENSO-
related phenomena
D.S. N° Regulation of environmental Art. The environmental manage-
004-2017- protection for the transporta- 33 ment plan must include mea-
MTC tion industry surements that regulate the
water use management, soil
and land use, wildlife and
fauna protection, erosion
control and reduction, and
treatment of emissions
D.S. N° Regulation of environmental Art. The construction of perfora-
039-2014- protection for activities related 78 tion platforms must include a
EM to hydrocarbons detailed study of slope stabil-
ity, erosion and sediment
control, and an evaluation of
the natural drainage systems
D.S. N° Regulation of environmental Art. The basic contents of the
019-2012- management in the agriculture 25 environmental impact report
AG industry must consider the water
usage, land cover conserva-
tion, soil degradation mea-
surements, biologic diversity
conservation, public health,
national heritage conserva-
tion, and basic service
infrastructure
D.S. N° Regulation of environmental Art. In the case of an existing
042-2017- protection for mining 17 nearby river or natural drain-
EM exploitation age line, erosion control
structures must be built
Boliviab Ley N° Environmental law Art. Mining operations must
1333 71 restore used areas, mitigate
erosion, assure slope stability,
and protect water streams,
during and after execution
(continued)
88 R. R. Gutierrez et al.

Table 2 (continued)
Country Codea Name Article Description
Art. Hydrocarbons-related activi-
73 ties must contain environ-
mental measurements of
prevention from deforesta-
tion, erosion, sedimentation,
and biota protection
a
“Ley” = Act, “D.S.” = Ordinance
b
Based on Ministerio de Medio Ambiente y Agua [139]

by Rosas and Gutierrez [7], global observations of agricultural lands extension for
the year 2000, and the extraction of soil degradation-related data from freely
available global datasets and model outputs available across these countries.
Our results indicate that soil erosion rates at agricultural lands have increased in
the last three decades in Peru and Bolivia. The largest areas of severely denuded soil
are persistently located in agricultural lands from the tropical rainforest, cold and hot
semi-arid, and tundra climate units in Peru and Bolivia. Likewise, although initial
and based on an observational criterion, our assessment reveals that there is a
positive relation between soil erosion and soil organic carbon stock. A further
analysis of this relationship is certainly needed.
Industrial-scale agriculture has markedly grown in Peru since the early 1990s,
especially in the Coastal section. In contrast, only flex crops have developed in
Bolivia in the last two decades. This agricultural growth dynamics appears to be
related to (1) the development of high soil salinity spots in the Peruvian Coast (e.g.,
Piura and Arequipa regions) and the Bolivian highlands (e.g., La Paz and the
surroundings of the Titicaca Lake) and (2) the increase in fertilizer pollution risk
in both countries, although this risk in Peru is one of the highest in Latin America
(only comparable to that in Ecuador and Mexico).
A limited number of local studies have addressed the presence of microplastics in
soils. They nevertheless report a sustained growing trend. The use of sewage-derived
waters of urban and peri-urban farms at large Peruvian cities and several areas of
Bolivia may result in high concentration microplastics spots in these locations.
Overgrazing affects a large portion of the highlands from both countries. Very
limited information regarding the extent of slash-and-burn agriculture is reported in
the scientific literature. However, it suggests that most of this practice is concen-
trated in the Department of Santa Cruz in Bolivia.
This contribution identifies some potential measures to limit the current high
pressure on soil resources in the study countries. Adopting new technology (e.g.,
remote sensors, UAV monitoring and mapping) and new agricultural techniques
(e.g., vertical farming, aeroponic cultivation) or the adoption of sustainable soil
management practices (e.g., zero tillage, responsible fertilizers use, and crop diver-
sification) appears to be feasible in the short/medium term.
In addition, establishing regulatory frameworks for erosion control in Peru and
Bolivia like the US National Pollutant Discharge Elimination System might limit the
Agricultural Land Degradation in Peru and Bolivia 89

increase of soil erosion rates. We contribute with a first digital assessment to


benchmark the current state of soil degradation in Peru and Bolivia.

Acknowledgments We acknowledge the Pontificia Universidad Católica del Perú (PUCP) for
supporting this contribution.

References

1. Khan OT (2014) Soil degradation, conservation and remediation. Springer


2. Karlen DL, Rice CW (2015) Soil degradation: will humankind ever learn? Sustainability 7:
12490–12501
3. Delang CO (2017) China’s soil pollution and degradation problems. Routledge
4. Qi Y, Yang X, Pelaez AM, Lwanga EH, Beriot N, Gertsen H, Garbeva P, Geissen V (2018)
Macro-and micro-plastics in soil-plant system: effects of plastic mulch film residues on wheat
(Triticum aestivum) growth. Sci Total Environ 645:1048–1056
5. Lal R (2013) Terrestrial carbon sequestration terrestrial. Reference module in earth systems
and environmental sciences. Elsevier
6. Borrelli P, Robinson DA, Fleischer LR, Lugato E, Ballabio C, Alewell C, Meusburger K,
Modugno S, Schütt B, Ferro V (2017) An assessment of the global impact of 21st century land
use change on soil erosion. Nat Commun 8:2013
7. Rosas MA, Gutierrez RR (2020) Assessing soil erosion risk at national scale in developing
countries: the technical challenges, a proposed methodology, and a case history. Sci Total
Environ 703:135474
8. Montaño-Lopez F, Guevara M, Biswas A (2021) Soil science research and development in
Latin America and the Caribbean. In: Rakshit A, Singh S, Abhilash P, Biswas A (eds) Soil
science: fundamentals to recent advances. Springer, Singapore, pp 613–621
9. OECD/FAO (2019) Latin American agriculture: prospects and challenges. In: OECD Pub-
lishing. OECD-FAO agricultural Outlook 2019-2028: Food and Agriculture Organization of
the United Nations, Rome, pp 70–124
10. Dasgupta P (2021) The economics of biodiversity: the Dasgupta review. HM Treasury,
London
11. Ananda J, Herath G (2003) Soil erosion in developing countries: a socio-economic appraisal. J
Environ Manag 68:343–353
12. Lal R (1998) Soil erosion impact on agronomic productivity and environment quality. Crit Rev
Plant Sci 17:319–464
13. Panagos P, Standardi G, Borrelli P, Lugato E, Montanarella L, Bosello F (2018) Cost of
agricultural productivity loss due to soil erosion in the European Union: from direct cost
evaluation approaches to the use of macroeconomic models. Land Degrad Dev 29:471–484
14. Berdegué JA, Fuentealba R (2011) Latin America: the state of smallholders in agriculture.
IFAD conference on new directions for smallholder agriculture
15. Urioste M (2012) Concentration and “foreignisation” of land in Bolivia. Can J Dev Stud 33:
439–457
16. Beckman J, Sands RD, Riddle AA, Lee T, Walloga JM (2017) International trade and
deforestation: potential policy effects via a global economic model
17. World Bank (2017) Gaining momentum in Peruvian agriculture: opportunities to increase
productivity and enhance competitiveness. World Bank
18. Martın-Retortillo M, Pinilla V, Velazco J, Willebald H (2019) The dynamics of Latin
American agricultural production growth, 1950–2008. J Lat Am Stud 51:573–605
90 R. R. Gutierrez et al.

19. Latrubesse EM, Restrepo JD (2014) Sediment yield along the Andes: continental budget,
regional variations, an comparisons with other basins from orogenic mountain belts. Geomor-
phology 216:225–233
20. Pepin E, Guyot J-L, Armijos E, Bazan H, Fraizy P, Moquet JS, Noriega L, Lavado W,
Pombosa R, Vauchel P (2013) Climatic control on eastern Andean denudation rates (Central
Cordillera from Ecuador to Bolivia). J S Am Earth Sci 44:85–93
21. Clark KE, Hilton RG, West AJ, Robles Caceres A, Gröcke DR, Marthews TR, Ferguson RI,
Asner GP, New M, Malhi Y (2017) Erosion of organic carbon from the Andes and its effects
on ecosystem carbon dioxide balance. J Geophys Res Biogeo 122:449–469
22. Tang FHM, Lenzen M, McBratney A, Maggi F (2021) Risk of pesticide pollution at the global
scale. Nat Geosci 14:206–210
23. Correa SW, Mello CR, Chou SC, Curi N, Norton LD (2016) Soil erosion risk associated with
climate change at Mantaro River basin, Peruvian Andes. Catena 147:110–124
24. Metternicht G, Zinck J (2016) Geomorphic landscape approach to mapping soil degradation
and hazard prediction in semi-arid environments: salinization in the Cochabamba Valleys,
Bolivia. In: Zinck JA, Metternicht G, Bocco G, Francisco Del Valle H (eds) Geopedology.
Springer, Cham. 425-439
25. NASADEM (2020) Merged DEM Global 1 arc second V001 [Dataset]. NASA EOSDIS Land
Processes DAAC
26. Vermote E, Justice C, Csiszar I, Eidenshink J, Myneni R, Baret F, Masuoka E, Wolfe R,
Claverie M (2014) NOAA Climate Data Record (CDR) of normalized Difference Vegetation
Index (NDVI), Version 4. NOAA Natl. Clim. Data Cent
27. Hengl T, Mendes de Jesus J, Heuvelink GBM, Ruiperez Gonzalez M, Kilibarda M,
Blagotić A, Shangguan W, Wright MN, Geng X, Bauer-Marschallinger B, Guevara M
(2017) SoilGrids250m: global gridded soil information based on machine learning. PLoS
One 12(2):1–40
28. Funk C, Peterson P, Landsfeld M, Pedreros D, Verdin J, Shukla S, Husak G, Rowland J,
Harrison L, Hoell A (2015) The climate hazards infrared precipitation with stations—a new
environmental record for monitoring extremes. Sci Data 2:1–21
29. Kinnell P (2005) Alternative approaches for determining the USLE-M slope length factor for
grid cells. Soil Sci Soc Am J 69(3):674–680
30. Williams JR, Berndt HD (1972) Sediment yield computed with universal equation. J Hydraul
Div 98:2087–2098
31. Ramankutty N, Evan AT, Monfreda C, Foley JA (2008) Farming the planet: 1. Geographic
distribution of global agricultural lands in the year 2000. Glob Biogeochem Cycles 22. https://
doi.org/10.1029/2007GB002952
32. Baied CA, Wheeler JC (1993) Evolution of high Andean Puna ecosystems: environment,
climate, and culture change over the last 12,000 years in the Central Andes. Mt Res Dev 13:
145–156
33. Pacheco P, Mejía E, Cano W, De Jong W (2016) Smallholder forestry in the Western Amazon:
outcomes from forest reforms and emerging policy perspectives. Forests 7:193
34. Piquer-Rodríguez M, Gasparri NI, Zarbá L, Aráoz E, Grau HR (2021) Land systems’
asymmetries across transnational ecoregions in South America. Sustain Sci 16:1519–1538
35. Molina LT, Gallardo L, Andrade M, Baumgardner D, Borbor-Córdova M, Bórquez R,
Casassa G, Cereceda-Balic F, Dawidowski L, Garreaud R (2015) Pollution and its impacts
on the South American cryosphere. Earth’s Future 3:345–369
36. Aguilar-Pesantes A, Peña Carpio E, Vitvar T, Koepke R, Menéndez-Aguado JM (2021) A
comparative study of mining control in Latin America. Mining 1:6–18
37. Peel MC, Finlayson BL, McMahon TA (2007) Updated world map of the Köppen-Geiger
climate classification. Hydrol Earth Syst Sci Discuss 4:439–473
38. FAO (2021) FAOSTAT Database. http://www.fao.org/faostat/
39. Tejada E, Arze ME, Moraes M, Bustillos F, Larrazabal DR, Trepp A, Leigue L, Mariscal CA,
Cabrera OJ, Gutierrez JM (2017) Food and nutrition security in Bolivia a country of
Agricultural Land Degradation in Peru and Bolivia 91

incalculable wealth. In: Clegg M, Bianchi E, McNeill J, Herrera L, Vammen K (eds)


Challenges and opportunities for food and nutrition security in the Americas. The View of
the Academies of Science, Mexico, The Inter-American Network of Academies of Sciences,
pp 52–75
40. Hofste R, Kuzma S, Walker S, Sutanudjaja E, Bierkens M, Kuijper M, Faneca Sanchez M,
Van Beek R, Wada Y, Galvis Rodríguez S, Reig P (2019) Aqueduct 3.0: updated decision-
relevant global water risk indicators. World Resources Institute
41. Shrivastava P, Kumar R (2014) Soil salinity: a serious environmental issue and plant growth
promoting bacteria as one of the tools for its alleviation. Saudi J Biol Sci 22(2):123–131
42. Wei X, Garcia-Chevesich P, Alejo F, García V, Martínez G, Daneshvar F, Bowling LC,
Gonzáles E, Krahenbuhl R, McCray JE (2021) Hydrologic analysis of an intensively irrigated
area in southern Peru using a crop-field scale framework. Water 13:318
43. Ivushkin K, Bartholomeus H, Bregt AK, Pulatov A, Kempen B, de Sousa L (2019) Global
mapping of soil salinity change. Remote Sens Environ 231:111260
44. Gamboa NR, Marchesem AB, Tavares Corrêa CH (2021) Salinization in Peruvian north coast
soils: case study in San Pedro de Lloc. In: Taleisnik E, Lavado RS (eds) Saline and alkaline
soils in Latin America. Springer, pp 141–159
45. Omuto C, Vargas R, El-Mobarak A, Mohamed N, Viatkin K, Yigini Y (2020) Mapping of salt-
affected soils. FAO, Rome
46. Lacroix P, Dehecq A, Taipe E (2020) Irrigation-triggered landslides in a Peruvian desert
caused by modern intensive farming. Nat Geosci 13(1):56–60
47. Olson K, Ebelhar S, Lang J (2014) Long-term effects of cover crops on crop yields, soil
organic carbon stocks and sequestration. Open J Soil Sci 4:284–292
48. Jia B, Niu Z, Wu Y, Kuzyakov Y, Li XG (2020) Waterlogging increases organic carbon
decomposition in grassland soils. Soil Biol Biochem 148:107927
49. Ruiz-Colmenero M, Bienes R, Eldridge D, Marques M (2013) Vegetation cover reduces
erosion and enhances soil organic carbon in a vineyard in the Central Spain. Catena 104:
153–160
50. Buytaert W, Cuesta-Camacho F, Tobón C (2011) Potential impacts of climate change on the
environmental services of humid tropical alpine regions. Glob Ecol Biogeogr 20(1):19–33
51. Vanacker V, Govers G, Poesen J, Deckers J, Dercon G, Loaiza G (2003) The impact of
environmental change on the intensity and spatial pattern of water erosion in a semi-arid
mountainous Andean environment. Catena 51(3):329–347
52. FAO (2002) Evaluación de la degradación de tierras en zonas áridas
53. Mendonça-Santos MdL (2015) Regional assessment of soil changes in Latin America and the
Caribbean. In: Status of the World’s Soil Resources (SWSR) – Main Report, pp 364–398
54. Sanderman J, Hengl T, Fiske GJ (2017) Soil carbon debt of 12,000 years of human land use.
Proc Natl Acad Sci 114(36):9575–9580
55. Gardi C, Angelini M, Barceló S, Comerma J, Cruz Gaistardo C, Encina Rojas AJA,
Krasilnikov P, Mendonça Santos Brefin M, Montanarella L, Muñiz Ugarte O, Schad P, Vara
Rodríguez M (2014) Atlas de suelos de América Latina y el Caribe. Dirección General de
Cooperación Internacional y Desarrollo (Comisión Europea), Luxembourg
56. Perrone A, Vassallo R, Lapenna V, Di Maio C (2008) Pore water pressures and slope stability:
a joint geophysical and geotechnical analysis. J Geophys Eng 5(3):323–337
57. Lim K, Li A-J, Cassidy M (2015) Pore pressure effect on slope stability assessment. ANZ
2015: changing the face of the earth - geomechanics and human influence. In: Proceedings of
the 12th Australia New Zealand conference on geomechanics, Wellington
58. Ke L, Takahashi A (2012) Strength reduction of cohesionless soil due to internal erosion
induced by one-dimensional upward seepage flow. Soils Found 52(4):698–711
59. Garcia-Chevesich P, Wei X, Ticona J, Martínez G, Zea J, García V, Alejo F, Zhang Y,
Flamme H, Graber A (2021) The impact of agricultural irrigation on landslide triggering: a
review from Chinese, English, and Spanish literature. Water 13:10
92 R. R. Gutierrez et al.

60. Torero M, Zegarra E, Pender J, Maruyama E, Keefe M, Stoorvogel J (2010) Socioeconomic


and technical considerations to mitigate land and water degradation in the Peruvian Andes.
CGIAR Challenge Program on Water and Food, Colombo
61. de Souza Machado AA, Lau CW, Till J, Kloas W, Lehmann A, Becker R, Rillig MC (2018)
Impacts of microplastics on the soil biophysical environment. Environ Sci Technol 52(17):
9656–9665
62. Zhu D, Chen Q-L, An X-L, Yang X-R, Christie P, Ke X, Wu L-H, Zhu Y-G (2018) Exposure
of soil collembolans to microplastics perturbs their gut microbiota and alters their isotopic
composition. Soil Biol Biochem 116:302–310
63. Purca S, Henostroza A (2017) Presencia de microplásticos en cuatro playas arenosas de Perú.
Rev Peru Biol 24(1):101–106
64. Manrique R (2019) Microplásticos sedimentos fluviales de la cuenca baja y desembocadura
del río Jequetepeque, Perú
65. Castañeta G, Gutiérrez AF, Nacaratte F, Manzano CA (2020) Microplásticos: Un
contaminante que crece en todas las esferas ambientales, sus características y posibles riesgos
para la salud pública por exposición. Revista Boliviana de Química 37(3):160–175
66. Horton AA, Walton A, Spurgeon DJ, Lahive E, Svendsen C (2017) Microplastics in fresh-
water and terrestrial environments: evaluating the current understanding to identify the
knowledge gaps and future research priorities. Sci Total Environ 586:127–141
67. Nizzetto L, Futter M, Langaas S (2016) Are agricultural soils dumps for microplastics of urban
origin? Environ Sci Technol 50(20):10777–10779
68. Thomas G (2014) Growing greener cities in Latin America and the Caribbean: an FAO report
on urban and peri-urban horticulture in the region
69. Echeverrıa I, Machicado L, Saavedra O, Escalera R, Heredia G, Montoya R (2019) Aguas
residuales domésticas tratadas con reactores anaeróbicos y filtros de grava como recurso para
ser usadas en agricultura. Investigación & Desarrollo 19:63–72
70. Butrón B (2020) Capacidad de biodegradación de Pseudomonas aeruginosa frente al
polietileno de baja densidad. Puno, Peru
71. Fremout T, Thomas E, Gaisberger H, Van Meerbeek K, Muenchow J, Briers S, Gutierrez-
Miranda CE, Marcelo-Peña JL, Kindt R, Atkinson R (2020) Mapping tree species vulnerabil-
ity to multiple threats as a guide to restoration and conservation of tropical dry forests. Glob
Chang Biol 26:3552–3568
72. World Bank (2007) Republic of Peru. Environmental sustainability: a key to poverty reduction
in Peru: country environmental analysis. World Bank
73. Lozada C (1991) Overgrazing and range degradation in the Peruvian Andes. Rangelands
Archives 13:64–66
74. Raboin ML, Posner JL (2012) Pine or pasture? Estimated costs and benefits of land use change
in the Peruvian Andes. Mt Res Dev 32:158–168
75. Polk MH, Young KR, Cano A, León B (2019) Vegetation of Andean wetlands (bofedales) in
Huascarán National Park, Peru. Mires & Peat 24:1–26
76. Swinton SM, Quiroz R (2003) Is poverty to blame for soil, pasture and forest degradation in
Peru’s Altiplano? World Dev 31:1903–1919
77. Montesinos-Tubée DB, Cleef AM, Sỳkora KV (2021) The subnival vegetation of Moquegua,
South Peru: chasmophytes, grasslands and cushion communities. Ecologies 2:71–111
78. Perevolotsky A (1991) Goats or scapegoats—the overgrazing controversy in Piura, Peru.
Small Rumin Res 6:199–215
79. LeBaron A, Bond LK, Aitke P, Michaelsen L (1979) An explanation of the Bolivian highlands
grazing-erosion syndrome. Rangeland Ecol Manag/J Range Manag Arch 32:201–208
80. Zimmerer KS (1993) Soil erosion and social (dis) courses in Cochabamba, Bolivia: perceiving
the nature of environmental degradation. Econ Geogr 69:312–327
81. MFA (1998) Evaluation of the Netherlands development programme with Bolivia. The
Netherlands
Agricultural Land Degradation in Peru and Bolivia 93

82. Muñoz MÁ, Faz Á (2014) Soil and vegetation seasonal changes in the grazing Andean
mountain grasslands. J Mt Sci 11:1123–1137
83. Mas-Coma S, Buchon P, Funatsu IRK, Angle R, Artigas P, Valero MA, Bargues Castelló MD
(2020) Sheep and cattle reservoirs in the highest human fascioliasis hyperendemic area:
experimental transmission capacity, field epidemiology and control within a one health
initiative in Bolivia. Front Vet Sci 7:841
84. Hartman BD, Bookhagen B, Chadwick OA (2016) The effects of check dams and other
erosion control structures on the restoration of Andean bofedal ecosystems. Restor Ecol 24:
761–772
85. Machaca NC, Condori B, Pardo AR, Anthelme F, Meneses RI, Weeda CE, Perotto-Baldivieso
HL (2018) Effects of grazing pressure on plant species composition and water presence on
bofedales in the Andes mountain range of Bolivia. Mires Peat 21:1–15
86. Fernandez J, Wickel B, Escobar M (2021) Hydro-ecological monitoring of high-elevation
wetlands in the Katari watershed, Bolivia. Sweden
87. Van Vliet N, Mertz O, Heinimann A, Langanke T, Pascual U, Schmook B, Adams C, Schmidt-
Vogt D, Messerli P, Leisz S (2012) Trends, drivers and impacts of changes in swidden
cultivation in tropical forest-agriculture frontiers: a global assessment. Glob Environ Chang
22:418–429
88. Naughton-Treves L, Mena JL, Treves A, Alvarez N, Radeloff VC (2003) Wildlife survival
beyond park boundaries: the impact of slash-and-burn agriculture and hunting on mammals in
Tambopata, Peru. Conserv Biol 17:1106–1117
89. White DC, Velarde SJ, Alegre JC, Tomich TP (2005) Alternatives to slash-and-burn (ASB) in
Peru, Summary Report and Synthesis of Phase II. Alternatives to Slash-and-Burn Programme.
Nairobi
90. Arce-Nazario JA (2007) Human landscapes have complex trajectories: reconstructing Peru-
vian Amazon landscape history from 1948 to 2005. Landsc Ecol 22:89–101
91. Smith J, Van De Kop P, Reategui K, Lombardi I, Sabogal C, Diaz A (1999) Dynamics of
secondary forests in slash-and-burn farming: interactions among land use types in the Peruvian
Amazon. Agric Ecosyst Environ 76:85–98
92. Pain A, Marquardt K, Lindh A, Hasselquist NJ (2021) What is secondary about secondary
tropical forest? Rethinking forest landscapes. Hum Ecol 49:239–247
93. Lindell L, Åström M, Öberg T (2010) Land-use change versus natural controls on stream water
chemistry in the Subandean Amazon, Peru. Appl Geochem 25:485–495
94. Killeen TJ, Guerra A, Calzada M, Correa L, Calderon V, Soria L, Quezada B, Steininger MK
(2008) Total historical land-use change in eastern Bolivia: who, where, when, and how much?
Ecol Soc 13:1–27
95. INE (2015) Censo Agroperucario 2013 Bolivia
96. Andersen LE, Doyle AS, del Granado S, Ledezma JC, Medinaceli A, Valdivia M, Weinhold D
(2016) Net carbon emissions from deforestation in Bolivia during 1990-2000 and 2000-2010:
results from a carbon bookkeeping model. PLoS One 11:e0151241
97. Bustillo Sánchez M, Tonini M, Mapelli A, Fiorucci P (2021) Spatial assessment of wildfires
susceptibility in Santa Cruz (Bolivia) using random forest. Geosciences 11:224
98. Wilkins JV (1991) The search for a viable alternative to slash and burn agriculture in the
lowland plains of Bolivia. Exp Agric 27:39–46
99. Kaimowitz D, Thiele G, Pacheco P (1999) The effects of structural adjustment on deforesta-
tion and forest degradation in lowland Bolivia. World Dev 27:505–520
100. Steininger MK (2000) Satellite estimation of tropical secondary forest above-ground biomass:
data from Brazil and Bolivia. Int J Remote Sens 21:1139–1157
101. Kennard DK (2002) Secondary forest succession in a tropical dry forest: patterns of develop-
ment across a 50-year chronosequence in lowland Bolivia. J Trop Ecol 18:53–66
102. Cisternas I, Velásquez I, Caro A, Rodríguez A (2020) Systematic literature review of
implementations of precision agriculture. Comput Electron Agric 176:105626
94 R. R. Gutierrez et al.

103. Mejía JC (2006) Agricultura de precisión en Bolivia. In: PROCISUR: Agricultura de preci-
sión: Integrando conocimientos para una agricultura moderna y sustentable
104. Instituto Nacional de Innovación Agraria (2018) MINAGRI aplica sensores remotos para
medir calidad de suelos agrícolas. https://www.inia.gob.pe/2018-nota-156/
105. de Althaus J (2008) La revolucion capitalista en el Perú. Fondo de Cultura Económica
106. Zhao J-C, Zhang J-F, Feng Y, Guo J-X (2010) The study and application of the IOT
technology in agriculture. In: 3rd International conference on computer science and informa-
tion technology, Chengdu
107. Wang Z, Zhao C, Zhang H, Fan H (2011) Real-time remote monitoring and warning system in
general agriculture environment. In: 2011 International conference of information technology,
computer engineering and management sciences, Nanjing
108. Oquelis Á, Landa D (2020) Desarrollo de un controlador agrícola para Agricultura de
Precisión con LoRaWAN para banano y mango orgánico
109. Colomina I, Molina P (2014) Unmanned aerial systems for photogrammetry and remote
sensing: a review. ISPRS J Photogramm Remote Sens 92:79–97
110. Kulbacki M, Segen J, Knieć W, Klempous R, Kluwak K, Nikodem J, Kulbacka J, Serester A
(2018) Survey of drones for agriculture automation from planting to harvest. In: 22nd IEEE
international conference on intelligent engineering systems, pp 353–358
111. Etyemezian V, Nikolich G, Nickling W, King J, Gillies J (2017) Analysis of an optical gate
device for measuring aeolian sand movement. Aeolian Res 24:65–79
112. De Clercq M, Vats A, Biel A (2018) Agriculture 4.0: the future of farming technology. World
Government Summit
113. Rose DC, Wheeler R, Winter M, Lobley M, Chivers C-A (2021) Agriculture 4.0: making it
work for people, production, and the planet. Land Use Policy 100:104933
114. Thomaier S, Specht K, Henckel D, Dierich A, Siebert R, Freisinger UB, Sawicka M (2015)
Farming in and on urban buildings: present practice and specific novelties of zero-acreage
farming (ZFarming). Renew Agric Food Syst 30(1):43–54
115. Kurumaa N (2021) Business out of vertical farming: is it possible?
116. Wittmann S, Jüttner I, Mempel H (2020) Indoor farming marjoram production—quality,
resource efficiency, and potential of application. Agronomy 10(11)
117. Eldridge BM, Manzoni LR, Graham CA, Rodgers B, Farmer JR, Dodd AN (2020) Getting to
the roots of aeroponic indoor farming. New Phytol 228(4):1183–1192
118. De Anda J, Shear H (2017) Potential of vertical hydroponic agriculture in Mexico. Sustain-
ability 9:140
119. Stein EW (2021) The transformative environmental effects large-scale indoor farming may
have on air, water, and soil. Air Soil Water Res 14:1–8
120. Scholl P, Leitner D, Kammerer G, Loiskandl W, Kaul H-P, Bodner G (2014) Root induced
changes of effective 1D hydraulic properties in a soil column. Plant Soil 381:193–213
121. Schawe M, Glatzel S, Gerold G (2007) Soil development along an altitudinal transect in a
Bolivian tropical montane rainforest: podzolization vs. hydromorphy. Catena 69:83–90
122. Günter S, Gonzalez P, Álvarez G, Aguirre N, Palomeque X, Haubrich F, Weber M (2009)
Determinants for successful reforestation of abandoned pastures in the Andes: soil conditions
and vegetation cover. For Ecol Manag 258:81–91
123. Jacobi J (2016) Agroforestry in Bolivia: opportunities and challenges in the context of food
security and food sovereignty. Environ Conserv 43:307
124. Walentowski H, Heinrichs S, Hohnwald S, Wiegand A, Heinen H, Thren M, Gamarra Torres
OA, Sabogal AB, Zerbe S (2018) Vegetation succession on degraded sites in the Pomacochas
Basin (Amazonas, N Peru)—ecological options for Forest restoration. Sustainability 10:609
125. Cotrina Sánchez DA, Barboza Castillo E, Rojas Briceño NB, Oliva M, Torres Guzman C,
Amasifuen Guerra CA, Bandopadhyay S (2020) Distribution models of timber species for
forest conservation and restoration in the Andean-Amazonian landscape, North of Peru.
Sustainability 12:7945
Agricultural Land Degradation in Peru and Bolivia 95

126. Brancalion PHS, Chazdon RL (2017) Beyond hectares: four principles to guide reforestation
in the context of tropical forest and landscape restoration. Restor Ecol 25(4):491–496
127. de Valença AW, Vanek SJ, Meza K, Ccanto R, Olivera E, Scurrah M, Lantinga EA, Fonte SJ
(2017) Land use as a driver of soil fertility and biodiversity across an agricultural landscape in
the Central Peruvian Andes. Ecol Appl 27(4):1138–1154
128. Strohmeier S, Fukai S, Haddad M, AlNsour M, Mudabber M, Akimoto K, Yamamoto S,
Evett S, Oweis T (2021) Rehabilitation of degraded rangelands in Jordan: the effects of
mechanized micro water harvesting on hill-slope scale soil water and vegetation dynamics. J
Arid Environ 185:104338
129. Clark KE, West AJ, Hilton RG, Asner GP, Quesada CA, Silman MR, Saatchi SS, Farfan-
Rios W, Martin RE, Horwath AB, Halladay K, New M, Malhi Y (2016) Storm-triggered
landslides in the Peruvian Andes and implications for topography, carbon cycles, and biodi-
versity. Earth Surf Dyn 4(1):47–70
130. Cremers L, Kolen J, de Theije MEM (2013) Small-scale gold mining in the Amazon. The
cases of Bolivia, Brazil, Colombia, Peru and Suriname. Cedla
131. Hidalgo VB, Dargent E (2020) State responses to the Gold Rush in the Andes (2004–2018):
the politics of state action (and inaction). Stud Comp Int Dev 55:516–537
132. Weisse MJ, Naughton-Treves LC (2016) Conservation beyond park boundaries: the impact of
buffer zones on deforestation and mining concessions in the Peruvian Amazon. Environ
Manag 58:297–311
133. Sears RR, Cronkleton P, Villanueva FP, Ruiz MM, del Arco MP-O (2018) Farm-forestry in
the Peruvian Amazon and the feasibility of its regulation through forest policy reform. Forest
Policy Econ 87:49–58
134. Román-Dañobeytia F, Huayllani M, Michi A, Ibarra F, Loayza-Muro R, Vázquez T,
Rodrıguez L, García M (2015) Reforestation with four native tree species after abandoned
gold mining in the Peruvian Amazon. Ecol Eng 85:39–46
135. Southgate D, Hitzhusen F, Macgregor R (1984) Remedying third world soil erosion problems.
Am J Agric Econ 66:879–884
136. Jha R, Whalley J (2001) The environmental regime in developing countries. In: Carraro C,
Metcalf G (eds) Behavioral and distributional effects of environmental policy. University of
Chicago Press, Chicago, pp 217–242
137. Swinton SM, Quiroz R (2003) Poverty and the deterioration of natural soil capital in the
Peruvian Altiplano. Environ Dev Sustain 5:477–490
138. Reyes K, Fernandez J, Llerena C, Cardozo J, Rojas G (2021) Perspectivas del estado de los
sedimentos en Perú. Perspectivas de la gestión actual de sedimentos en nueve países de las
Américas:107–119
139. Ministerio de Medio Ambiente y Agua (2014) Programa plurianual de gestión integrada de
recursos hídricos y manejo integral de cuencas. Bolivia
140. Pitt R, Clark SE, Lake DW (2007) Construction site erosion and sediment controls: planning,
design and performance. DEStech Publications, Inc
141. FHWA (2016) Construction program guide - erosion and sediment control
142. Rivas T (2006) Erosion control treatment selection guide
143. Cooke SJ, Chapman JM, Vermaire JC (2015) On the apparent failure of silt fences to protect
freshwater ecosystems from sedimentation: a call for improvements in science, technology,
training and compliance monitoring. J Environ Manag 164:67–73
Agricultural Land Degradation in Brazil

Paulo Tarso S. Oliveira , Raquel de Faria Godoi ,


Carina Barbosa Colman , Jaíza Santos Motta , Jullian S. Sone ,
and André Almagro

Contents
1 Insights from Agriculture and Land Degradation in Brazil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2 Water Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
2.1 Cerrado . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
2.2 Atlantic Forest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.3 Amazon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.4 Pantanal, Pampa, and Caatinga . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.5 Water Erosion Studies in Brazil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
2.6 An Estimate of Water Erosion Rates in Brazil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
2.7 Soil and Water Conservation in Brazil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3 Wind Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4 Slash-and-Burn Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5 Soil Compaction and Overgrazing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6 Salinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
7 Agrochemicals Use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
8 Microplastics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
9 Challenges and Future Perspectives for Agriculture and Land Degradation in Brazil . . . . 115
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

P. T. S. Oliveira (✉)
Faculty of Engineering, Architecture and Urbanism and Geography, Federal University of Mato
Grosso do Sul, Campo Grande, MS, Brazil
Department of Hydraulics and Sanitation, São Carlos School of Engineering, University of São
Paulo, São Carlos, SP, Brazil
R. de Faria Godoi, C. B. Colman, J. S. Motta, and A. Almagro
Faculty of Engineering, Architecture and Urbanism and Geography, Federal University of Mato
Grosso do Sul, Campo Grande, MS, Brazil
J. S. Sone
Department of Hydraulics and Sanitation, São Carlos School of Engineering, University of São
Paulo, São Carlos, SP, Brazil

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 97
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 97–128, DOI 10.1007/698_2022_923,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 14 December 2022
98 P. T. S. Oliveira et al.

Abstract In the last 50 years, Brazil changed its position of food importer to
become one of the main global players in food, fiber, and renewable energy pro-
ductions. This shift was only possible because of the country’s favorable climate
condition, soil types, relief, and environmental diversity, together with agricultural
technological advances such as the use of no-till, level terraces, development of
adapted seeds, intensification of mechanized agriculture, and currently smart agri-
culture. However, the country still faces problems with agricultural land degradation,
such as soil erosion (water and wind), slash-and-burn agriculture, soil compaction
and overgrazing, salinity, the abusive use of agrochemicals, and microplastics. In
this chapter, we discuss some of these issues on the agricultural land degradation in
Brazil, showing the main challenges, advances, and perspectives on this topic.

Keywords Food security, No-till, Rotational grazing systems, Soil erosion,


Soybeans

1 Insights from Agriculture and Land Degradation


in Brazil

Brazil is a country with continental dimensions (total area of ~8,500,000 km2), home
to a great diversity of landscapes, a large range of elevations (sea level to 2,900 m),
different hydrometeorological conditions, and a variety of rich ecosystems. The
country has six biomes – the Amazon, Atlantic Forest, Caatinga, Cerrado, Pantanal,
and Pampa [1]. The country is also known currently as one of the main global
players in agriculture and beef production. Agribusiness has been the main sector of
the Brazilian economy, representing one-third of Brazil’s gross domestic product
(GDP). For instance, in 2020 the Brazilian GDP was R$ 7.45 trillion and the
agribusiness GDB of R$ 2.0 trillion [2]. However, this was not the reality 50 years
ago, when the country was a food importer.
During the 1960s, only 2% of the Brazilian farmers had agricultural machinery
[3]. In general, soil management began with the elimination of undisturbed vegeta-
tion, followed by soil tillage and sowing performed using animal traction, and ended
with the burning of residues after harvesting. In the 1960s and 1970s, Brazil
experienced rapid urbanization and middle-class expansion, increasing the purchas-
ing power of households and changing the food consumption patterns. Domestic
food production was insufficient thanks to poor agricultural technology and land
management, while food imports were becoming unsustainable. The first page of a
national newspaper on April 24, 1968, was “The food scarcity in Brazil,” showing
the issues and challenges that the country was going through.
In the 1970s, Brazilian agriculture had the first steps toward more efficient
production. In that decade, the soil was managed with more intensity, using plows
and grades (plowing, chipping, and leveling), and sowing with mechanical traction,
using limestone to correct soil acidity and phosphate fertilizers. In general, crop
Agricultural Land Degradation in Brazil 99

Fig. 1 A summary of the agricultural development in Brazil

residues were still burned after harvesting. We also need to highlight the initial use of
terrace farming and contour plowing [4]. In addition, on December 7, 1972, the
Brazilian president Emílio Garrastazu Médici created the Brazilian Agricultural
Research Corporation (in Portuguese Empresa Brasileira de Pesquisa Agropecuária –
Embrapa). The Embrapa was of such importance to improve agriculture efficient
production, bringing, together with the universities, science and technology to the
Brazilian farmers.
After the 1970s, year by year the Brazilian agriculture has proved to be robust,
efficient, and extremely productive. The traditional soil management system that
used plows and harrows (conventional tillage) was replaced by a system involving
soil cover with crop residues and reduction/elimination of mechanical tillage
(no-till). With the use of direct seeding, water erosion and crop costs were reduced
in relation to conventional tillage. Currently, the country is one of the world’s largest
producers and exporters of grain and beef [1]. Figure 1 shows the development of
Brazilian agriculture divided into three steps: the 1960s, 1970s, and currently.
It has been reported that food production needs to increase by 60% to feed the
world population in 2050 [5]. Pastor et al. [6] have shown that the global agricultural
expansion into undisturbed lands may be around 300 Mha (100 Mha cropland and
200 Mha grassland) between 2000 and 2050. This projection was done considering
an increase in food demand, with more than 60% of the increase coming at the
expense of forests. Furthermore, 50–60% of this total conversion would occur in
Latin America according to the authors. This farmland expansion has occurred
mainly in the Brazilian Cerrado in one of the largest expanding agricultural frontiers
in the world known as MATOPIBA [1]. Therefore, it is clear that the Brazilian
agribusiness has been crucial for global food trade. Land-use and land cover
planning supported by new technologies is key to keep improving food, energy,
and fiber production in Brazil. Furthermore, it is also needed to solve important
issues related to soil degradation, such as soil erosion, slash-and-burn agriculture,
soil compaction and overgrazing, salinity, the abusive use of agrochemicals, and
microplastics. Some of these issues are discussed below in this chapter.
100 P. T. S. Oliveira et al.

2 Water Erosion

Water erosion is a prominent threat to Brazilian soils. The removal of the fertile
topsoil reduces crop productivity and produces polluting sediments that will be
transported downstream and may cause siltation and flooding. Water erosion in
Brazil is mainly related to its historical intensive agriculture expansion with a lack
of agricultural technology and unplanned occupancy. Brazil’s agriculture expansion
progressed mainly based on forest conversion to pasture and crops under manage-
ment systems that were inadequate to the pedoclimatic and topographic conditions
[7–9]. However, currently new agricultural technologies and soil and water conser-
vation practices have supported Brazilian farmers to improve productivity without
increasing soil erosion [10].
The high land availability, abundant water resources, and climatic conditions
have allowed Brazil to intensively expand its agricultural production while
protecting 67% of its native vegetation (Fig. 2) [11]. Even facing adversity in
managing its agricultural soils and using a small fraction of its land when compared
to other countries (i.e., 7.5% for crops and 19.7% for grazing), Brazil is still a major
food producer and exporter. The country is the world’s leading producer of soybean,
sugarcane, and coffee, the second major producer of beef cattle, and the third major
producer of chicken meat [12].
In a scenario where the conservation of remaining forest areas is sought, to
enhance food production, the amount of food produced per unit of land must
increase instead of increasing the agricultural area. Studies have shown that Brazil
has the unique opportunity to enhance agricultural production to meet projected food
demand at the same time conserving its biodiversity [13–15]. The balance between
soil protection, greenhouse gas emissions, and food security in an agriculture
intensification plan has been discussed [16–19]. Although the strategies for intensi-
fication have yet to be settled, one thing is sure: to effectively guide sustainable
agriculture intensification, knowledge on soil resources and the related degradation
processes is essential. Thus, the current status of agricultural soils in Brazil
concerning water erosion is briefly discussed here in more detail for each biome.

2.1 Cerrado

The land-use intensification in the Cerrado is one of the main causes of the current
powerful agri-food production of Brazil. For instance, from the late 1970s to 2020 in
the central-western region (where the Cerrado is the predominant biome), land use
for grain production increased 6.7 times, while the production increased almost
23 times [20]. Being a region with naturally low fertility soils, this growth was
only possible because of agricultural research on genetic improvement and acidity
correction, besides intense use of fertilizers [3, 21].
Agricultural Land Degradation in Brazil 101

Fig. 2 Land use and land cover summary of Brazil, depicting the coverage of the nine most
common classes in each biome. Forests alone correspond to almost half of Brazil’s land. Agricul-
tural land occupies 30% of the country’s territory

Currently, 41% of the country’s land use for agriculture and 36% for livestock
take place in the Cerrado (Fig. 2). The intensive change in land use in the biome
leads to land degradation, as it has been shown that the crop productivity loss more
than doubled in areas under severe erosion between 2000 and 2012, reaching an
annual soil loss rate of 12.0 Mg ha-1 year-1 [22]. The Cerrado portion in the states
of Maranhão, Tocantins, Piauí, and Bahia (MATOPIBA) is the newest agricultural
hotspot: the area for crops in this region more than doubled in the last two decades,
coming mostly from the conversion of natural vegetation [23, 24]. Clearing
Cerrado’s native vegetation increases soil loss rates in orders of magnitude and
consequently decreases agricultural productivity [22, 25, 26].
102 P. T. S. Oliveira et al.

2.2 Atlantic Forest

Concerning the forest biomes, the Atlantic Forest concentrates 32 and 18% of
Brazil’s land for agriculture and pasture, respectively, and 71% of lands of mosaic
agriculture (Fig. 2). The major part of Brazil’s forest plantations and sugarcane crops
is located in the Atlantic Forest. The latter were planted mostly over areas of
degraded pastureland and cropland [27], where soils have faced severe erosion
with rilling and gullying for many years [28, 29]. No-till with straw mulching has
been increasingly used in sugarcane crops to avoid accelerated erosion, although soil
compaction caused by the intense machinery traffic still poses a major limitation to
the spread of no-till in sugarcane crops [30]. Legal and political efforts associated
with market mechanisms have made the Atlantic Forest move toward reducing
deforestation rates and enhancing forest transition [31–33]. However, soil data
have been neglected in restoration projects [34]. Land management in the Atlantic
Forest requires major improvements concerning soil conservation [27].

2.3 Amazon

Cattle ranching and soybean crops have been put as the major causes of forest
clearing in the Amazon [35, 36]. The Amazon comprises almost one-third of Brazil’s
pastureland (Fig. 2). Although these areas are the outcome of deforestation, grazing
activity in already-cleared areas has been considered sustainable use of Amazon land
[37, 38]. Degraded pastures are responsible for 50% of total soil loss caused by
agricultural activities in Brazil [39]. In the Amazon, degraded pastureland reduced
more than nine times between 2008 and 2014 [40]. This fact is a response to
environmental policies such as the Soy Moratorium that incentivized soy expansion
in already-cleared areas, prioritizing degraded pastures [24].
Water erosion rates in soybean crops are diminished as at least half of the crops
are planted under no-till cultivation [39]. However, the market’s pressure for crop-
land expansion creates off-site effects, boosting the risk of deforestation advancing
north [24, 36, 41]. Wildfires in the Amazon also increase the risk of water erosion
[42], especially in lowland forests where lower tree cover densities occur [43] and
soil erodibility [44] and rainfall erosivity are high [45, 46]. If not controlled, soil
erosion can change Amazon rainforests into a savanna state [47].

2.4 Pantanal, Pampa, and Caatinga

The Pantanal, Pampa, and Caatinga biomes also offer many opportunities for
sustainable use, for which land-use planning and conservation policies are essential
[48, 49]. For instance, in the Pampa biome, the natural grasslands offer shelter for
Agricultural Land Degradation in Brazil 103

sustainable large-scale livestock grazing [50, 51], but previous inadequate manage-
ment practices and financial interests for quick profit made farmers trade the
grasslands’ potential for cattle production for other agriculture activities [52]. In
fact, agriculture currently represents 41% of the biome’s area (Fig. 2), being the main
rice and wheat producer in Brazil [53]. Once the soils in the Pampa biome are very
susceptible to water and wind erosion, inappropriate land use represents a major
threat to this fragile environment [54].
In the Pantanal, ranchers have been conciliating low-density cattle grazing with
landscape conservation for two centuries and a half [55]. However, increasing
pressure coming from intense land use in the surrounding highlands (Cerrado and
Amazon biomes) represents a major threat to the Pantanal ecosystem [56]. The
lowlands and the highlands dynamics are strictly connected, so that the Cerrado
rivers carry the increased sediment load caused by water erosion to the deposition
plain, worsening already existing siltation problems in the Pantanal [57] and threat-
ening the sustainable management in the region.
The Caatinga biome is a semi-arid region located in the Northeast of Brazil, being
the driest region of the country. It comprises mostly secondary vegetation (herba-
ceous and arboreous) and presents a severe dry season, where ~70% of the annual
precipitation occurs between February and April. Despite this dry condition, the
lands of the Caatinga are destined for grazing livestock on pasture (25.2%) and
agriculture production (11.3%) by using irrigation and rain-fed agriculture (Fig. 2).
Most of the farmlands in this region are rain fed, although irrigated areas have
significantly increased in the last few years [58]. Furthermore, there is a concentra-
tion of small family farmers in this region. There is potential to increase food
production in this region by using irrigation; therefore, soil and water conservation
techniques are necessary to improve productivity and to control the soil erosion
effects.

2.5 Water Erosion Studies in Brazil

The number of soil erosion studies with field experiments in Brazil has decreased by
86% since 2000 [59]. Although direct measurements are essential for the modeling
validation process, indirect methods to predict soil erosion by water have been
extensively and increasingly applied in Brazil. The empirical Universal Soil Loss
Equation (USLE) by Wischmeier and Smith [60] is the most widespread model to
predict soil loss by water in Brazil [61]. The recent advances in remote sensing and
environmental data availability allow researchers to approach the USLE six envi-
ronmental risk factors in large-scale studies.
Almagro et al. [46] have estimated the rainfall erosivity (R) and its projected
changes across Brazil. The study finds increased projections of R-factor in north-
eastern and southern Brazil, and a decrease of as much as -71% in the 2071–2099
period was estimated for the southeastern, central, and northwestern parts of the
country. Knowing the change patterns in rainfall distribution is essential for planning
104 P. T. S. Oliveira et al.

soil and water conservation. The soil erodibility factor (K) has been indirectly
determined by Godoi et al. [44] based on the USLE nomograph and validated against
field measurements. The study identifies critical erodibility values for the Western
Amazon soils, where the forests present a crucial role in protecting against water
erosion. The results provide support for the sustainable management of soil
resources.
The improvements in terrain data and algorithms allow a more precise estimation
of the topographic factor (LS). For instance, the LS tool developed by Zhang et al.
[62, 63] is an improved method to estimate the slope length and USLE LS factor.
This method presents an approach for overcoming the limitations of slope length
determination based on flow path and cumulative cell length-based method (FCL).
Refined digital elevation models (DEMs) have been developed such as the MERIT
DEM created by Yamazaki et al. [64]. The improved DEM removes major error
components from existing digital elevation models (DEMs), providing substantial
improvements for flat areas, such as the Amazon floodplain.
With the decline of plot-scale erosion studies, the most recently applied methods
for the determination of the cover-management factor (C) are based on geographic
information system (GIS) procedures, basically by the assignment of literature
C-factor values to land-use maps or the application of vegetation indices. Although
these methods may not adequately represent the management aspect of the C-factor,
they allow a better understanding of the soil erosion potential in large-scale appli-
cations [65]. Durigon et al. [66] propose a method for calculating C-factor based on
the normalized difference vegetation index (NDVI) for tropical regions. This method
was investigated by Almagro et al. [67] against literature values and was considered
the most suitable alternative for computing soil loss in tropical regions. Most
recently, Macedo et al. [68] proposed a new approach for computing the C-factor
considering the effect of low-reflectance vegetation cover areas on the reduction of
the effects on erosion caused by rainfall seasonality. The authors found a better
C-factor estimating by their approach than that proposed by Durigon et al. [66]; thus,
this is also a suitable approach to be used in Brazil.

2.6 An Estimate of Water Erosion Rates in Brazil

Combining the above-discussed methods and data for R [46], K [44], LS [62–64],
and C [66] factors, we have the necessary inputs to estimate the average soil loss rate
by water erosion using the USLE, by simply multiplying the factors. To estimate the
C-factor, we used MODIS imagery of the last 10 years (2011–2021) as input to the
formula proposed by Durigon et al. [66], with the adjustment factor of 0.1 reported
by Colman [69]. In the first moment, the support practice factor (P) was not
evaluated. The resulting map of water erosion for Brazil is presented in Fig. 3.
According to the present mapping, approximately 81.7 Mha or 9.6% of Brazil’s
land is under severe erosion (>11 t ha-1 year-1). Considering the arable land (areas
currently under agricultural use and grasslands), 28.4 Mha or 9.4% are severely
Agricultural Land Degradation in Brazil 105

Fig. 3 Map of soil loss by water in Brazil. The numbers outlined by the dashed border indicate the
biomes: 1 – Amazon, 2 – Caatinga, 3 – Cerrado, 4 – Pampa, 5 – Pantanal, and 6 – Atlantic Forest

eroded. In terms of biomes, the Atlantic Forest is the most threatened one with
almost 25% under severe erosion, followed by the Cerrado with 9%. Figure 4
presents the coverage of each erosion interval for each land-use class. Non-forest
ecosystems (NFE; here basically composed of grasslands) is the class most affected
by water erosion, while wetlands represent the least susceptible one.

2.7 Soil and Water Conservation in Brazil

The benefits of conservation agriculture on water erosion control are known by


Brazilian farmers since the 1970s with the application of no-tillage systems (NTS),
but only from the 1990s the system was established, experiencing a rapid expansion
106 P. T. S. Oliveira et al.

Fig. 4 The stacked bars indicate the coverage of each soil loss interval for the main land-use classes
in Brazil

to the present day achieving ~33 Mha [70–72]. NTS performed a crucial role in the
progress of Brazilian agriculture, expanding its double-cropping area by reducing
the soil tillage [3]. However, the system that is based on minimal soil disturbance,
permanent ground cover, and crop rotation has been “simplified” by farmers with
harmful techniques such as monoculture continuous cropping, straight-row farming
up and down the hill, and even eliminating terraces [73].
Soil research institutions have been gathering forces to rescue the principles of
conservation agriculture and improve farming techniques in Brazil. It is estimated
that NTS is applied in 86% of the area of soybean, maize, and bean crops [3]. If
correctly applied, no-tillage systems can reduce soil losses by up to 90% as com-
pared with the conventional tillage [74]. Integrated crop–livestock and crop–
livestock–forest systems (ICLFS) are also a reality in 11.5 Mha in Brazil, where
the improvement of soil quality with increased C accumulation, better soil resilience,
and increased crop and meat productivity have been reported [75, 76]. Integrated
systems under no-tillage offer an opportunity to recover and produce on sandy soils
previously considered inhospitable to agricultural production [77].

3 Wind Erosion

Wind erosion is a movement of soil forced by the wind that occurs in dry conditions
when the soil is exposed to the wind [78]. In the face of climate change, higher
temperatures, reduced rainfall, and changes in wind speed are expected [79]. While
the increase in the wind speed may increase wind power generation [80], the wind
can become an intensifier of soil loss, especially when the soil is uncovered,
unstructured, or not in a natural condition. With changes in land use and land
Agricultural Land Degradation in Brazil 107

cover associated with climate change, wind erosion can cause major social, envi-
ronmental, and economic impacts [81].
Although studies on water erosion are more common (as seen in the topic water
erosion) in Brazil, wind erosion can happen in certain situations and in places with
greater wind occurrence and where the soil is more exposed, such as in the regions
from the northeast of Brazil and in the state of Rio Grande do Sul [82, 83]. However,
studies on the impacts of wind erosion in Brazil have received considerably less
attention, mainly due to the difficulty of observation, the low amount of field
experiments, and the low availability of data. A study carried out on a plot-scale
study in Rio Grande do Sul state is a closer example of wind erosion assessment, in
which the movement of sand caused by the wind was evaluated [84].
The modeling of wind erosion can be done using equations that consider factors
such as wind speed, soil, vegetation, and land use and management [81]. The field
erosion model revised wind erosion equation (RWEQ) [78] is one of the most
applied equations in wind erosion modeling in the world [82, 85, 86]. In general,
the input data needed to run the model are usually measured directly [82]. In Brazil,
although climatological and hydrological data are monitored over time [87, 88], the
country lacks field experiments to validate the model. In addition, research groups
need to be equipped with a high computational power to perform the modeling and
with a large data storage capacity to run the entire continental extent of Brazil.
The increase in the rates of soil loss due to wind is a challenge on site that reduces
the capacity of soils to support vegetation and livestock [89]. At the same time, it
also causes impacts off site the farm related to the spread of dust, herbicides, and
pesticides [89]. Therefore, knowledge of wind erosion is essential for the mainte-
nance of soils in Brazil. The potential for soil loss due to wind erosion is still little or
not known in Brazil. It remains unclear where, when, and to what extent wind
erosion can affect Brazilian arable land. Therefore, field observations and technical
and financial support are required. We encourage researchers about the importance
of local measurements, observations, and future research on a field scale, seeking a
better understanding of the complex issue of the dynamics of wind erosion for better
use and management of soil in Brazil.

4 Slash-and-Burn Agriculture

Slash and burn is one of the most primitive and employed agricultural techniques in
the world, especially in tropical areas. It comprises the slash of the native vegetation
and later burn of the biomass to convert the area to permanent cropland or pasture.
Due to the low cost, lack of machinery, and governmental incentives, this traditional
agricultural method is still very much utilized by smallholders and farmers.
In Brazil, slash-and-burn agriculture dates to centuries ago, by indigenous,
caboclos, caiçaras, and quilombolas communities, which adopted the slash and
burn to their subsistence [90–93]. It is important to note, as stated by Serrão et al.
[92], that there is a huge difference between the slash-and-burn agriculture employed
108 P. T. S. Oliveira et al.

by these traditional Brazilian communities and the small farmers who migrated to
Brazil. Traditional communities’ slash and burn can be considered sustainable due to
their ecological knowledge acquired through ages of farming and low cropland
density. On the other hand, the migrant farmers are often susceptible to a quick
move for a new area after a slash-and-burn cycle, due to a too short (or inexistence
of) fallow period, leading to excessive deforestation in a “never-ending” cycle of
deforestation.
Nowadays, the employment of slash-and-burn agriculture is mainly concentrated
in the Amazon, by smallholders for their supply. However, there are also large areas
where the slash-and-burn technique is employed to convert into pasture for live-
stock. In 1996, it was estimated that more than 500,000 farmers were used to
applying the slash-and-burn technique throughout the Amazon, especially at the
“Arc of Deforestation” [92]. Given the actual situation of the Amazon, being one of
the most prominent agricultural frontiers in Brazil [94], it is expected that this
number of farmers adopting the slash-and-burn technique is highly underestimated.
Slash-and-burn agriculture in southern Brazil has different characteristics than
that employed in forest areas. In this region, the technique is applied in well-
established agricultural landscapes, mainly to black beans and maize cultivation
[95]. The first Italian settlers and immigrants also employed the technique in
southern Brazil but also fertilized the soil with rotating crops. They implemented
mainly familiar agriculture with mixed crops from Brazil and Italy, such as tobacco,
potatoes, wheat, rice, maize, and beans [96].
There are many negative impacts related to the slash and burn, such as high
greenhouse gas emissions, loss of biodiversity and soil nutrients, deforestation, and
accidents related to smoke, aside from the risk of uncontrolled forest fires. In a study
case in the Caatinga biome, Mamede and de Araújo [97] suggest that successive
slash-and-burn events with short fallow intervals may contribute to diminishing
livestock carrying capacity, by lowering native plants species diversity.
Despite the ease of application, slash-and-burn agriculture does not have the
capacity for the needs of growing demand and sustainable development
[98]. Then, many techniques have been suggested as alternatives to slash-and-burn
(ASB) agriculture to ensure the ecosystem services provisioning and well-being of
local communities. Comte et al. [98] suggest that by permitting fallows in short
intervals between slash-and-burn implementation, and an increased period of culti-
vation to an amelioration of the soil quality, it is possible to promote more sustain-
able intensification of agriculture in Amazon. Tomich et al. [99] and Tremblay et al.
[100] suggest that agroforestry with medium- and high-size tree crops is a viable
alternative to local farmers to abolish slash-and-burn agriculture in the Amazon, with
profitable results from the farmer’s perspective and sustainable management from
the ecological perspective. Even the implementation of improved pastures could be
profitable for local farmers, while this would not achieve public global concerns of
carbon stocks, biodiversity, and greenhouse gas absorption [99].
In all the abovementioned studies, there is a common belief that all the alterna-
tives to the slash-and-burn techniques are dependent on policy-makers’ funding and
incentives to the local farmers, lowering the risks for the communities and nature. In
Agricultural Land Degradation in Brazil 109

this way, the implementation of a standard framework and technology development


is essential.

5 Soil Compaction and Overgrazing

Agricultural intensification plays a key role in meeting the ever-increasing global


demand for food. Nonetheless, intensification of production becomes critical when
environmental aspects are neglected, leading to a disturbance of soil ecosystem
functioning and a depletion of its capacity of sustaining food production. Intensive
use of farm machinery and heavier livestock stocking rates has led to soil compac-
tion and deterioration of soil physical properties. Overall, 33% of the world’s soils
are moderately to highly degraded due to erosion and compaction [101]. In Brazil,
for instance, agricultural lands have an annual absolute land productivity loss of
6.4% due to soil degradation triggering a decrease of 385 thousand tons in livestock
production [102]. This current scenario indicates that soil degradation clearly leads
to substantial economic costs and losses.
Pasturelands cover about 19.7% of the Brazilian territory (~1,670,000 km2;
Fig. 2) and are the main sustenance for ~253 million cattle head, which allowed
the country to become the largest beef producer and exporter in the world
[103]. Livestock production occurs mainly in the Amazon, Cerrado, and Pantanal
biomes (Fig. 5). Nevertheless, about 50% of the pasture area presents some level of
degradation compromising the productivity, which is, according to Strassburg et al.
[13], at 32–34% of its potential. Overgrazing is the main driver of soil degradation

Fig. 5 Cattle herd (a) and cattle density (b) by municipality in the Brazilian biomes: 1 – Amazon,
2 – Caatinga, 3 – Cerrado, 4 – Pampa, 5 – Pantanal, and 6 – Atlantic Forest. Cattle herd data were
obtained from IBGE’s Census of Agriculture [53]. To estimate the cattle density per municipality,
the number of heads was divided by the pasture area mapped by MapBiomas [11]. For Pampa and
Pantanal biomes, native grassland area was added to pasture area
110 P. T. S. Oliveira et al.

and compaction due to cattle trampling. This negative impact on the soil physical
properties increases soil bulk density and penetration resistance and decreases soil
porosity and water infiltration. Thus, management practices are key to increase
production through agricultural intensification while mitigating the deterioration of
soil physical properties.
Sone et al. [104] showed that site-specific management practices can increase
livestock productivity and provide higher water infiltration into the soil. They
investigated the impacts of intensive grazing management on water infiltration and
found a negative correlation between grass cover and soil bulk density (i.e., a pasture
with an optimal grazing pressure maintains an adequate grass cover and height,
which prevents the soil from compaction). Moreover, increasing productivity by
49–52% of the Brazilian pastures’ potential would be sufficient to meet the demand
for beef, crops, wood, and biofuels until at least 2040 without further conversion of
natural ecosystems [13]. It shows the potential of the Brazilian agriculture to increase
production through intensification while sparing land and avoiding soil
deterioration.
Alternatives to avoid soil compaction by overgrazing are the adoption of site-
specific optimal grazing pressures along with adaptable stocking rates and rotational
grazing [104–106]. There is no general rule of thumb for universal grazing pressure
or stocking rate to maintain the long-term sustainability of beef production systems
in Brazil. Another alternative to avoid soil degradation is the use of fertilizers to
encourage forage growth and, consequently, make the soil more resistant to tram-
pling [104, 107]. However, fertilization should always be done in the proper amount
and at the right time, taking into account the characteristics of the soil and climate of
the area, associating it with good management practices [108, 109]. Pasture restora-
tion and maintenance costs can be offset by the sustainable intensification of grazing
that can provide higher animal and land productivity and, in turn, increase the
profitability of the production chain. For instance, in Brazil, the adoption of forage
grass cultivars and legumes in about 41 million ha generated an annual net benefit of
US$ 3.45 billion in 2012 [14, 110].
Croplands comprise around 7.5% of the Brazilian total land area (~640,000 km2;
Fig. 2) [11, 111]. Intensive machinery traffic compacts the first layers of soil
increasing the soil vulnerability to soil erosion by water, which removes the
nutrient-rich topsoil and consequently decreases crop yield. Crop residue and
no-till management have been feasible alternatives for coping with these challenges
in agriculture worldwide. No-tillage systems are mainly based on crop rotation and
the use of cover crops and crop residues for maintaining a continuous soil cover.
More than 320,000 km2 of the Brazilian croplands are under no-till [3], and it has
contributed to addressing the devastating soil erosion caused by intense tillage in the
1970s – when no-till farming started in Brazil. One of the disadvantages is that the
long-term no-tillage practice may increase soil bulk density and soil penetration
resistance [112], which is minimized by adopting periodic subsoiling [113].
Currently, no-till farming is generally known as conservation agriculture and –
along with adequate crop, nutrient, weed, and water management – is the basis of the
sustainable agricultural intensification strategy developed by the Food and
Agricultural Land Degradation in Brazil 111

Agriculture Organization (FAO). The pressure on agricultural intensification in


Brazil led to an increase in machinery traffic, exacerbating soil compaction. In the
Cerrado and Pampa biomes, the main sources of shallow soil compaction and poor
structural conditions of the topsoil are also associated with soybean monoculture and
long fallow periods during winter [114, 115]. This practice does not comply with
guidelines of the conservation agriculture. Therefore, it is of paramount importance
to improve productivity in an environmental, social, and economical way. Moreover,
these characteristics corroborate the climate-smart agriculture, which aims at
increasing productivity, strengthening farmers’ resilience to climate change, reduc-
ing GHG emissions, and sequestering carbon [116]. In a Brazilian context, Sone
et al. [10] found an improvement of water infiltration and a reduction of soil erosion
due to no-till farming, crop diversification, crop–livestock rotation, and crop–
livestock–forestry integration, compared with conventional agriculture. They
observed that water infiltration and soil loss rates from those agricultural systems
were similar to native Cerrado vegetation (i.e., Cerrado sensu stricto). This evinces
the great potential of Brazil for meeting the future global food demand by adopting
the FAO’s sustainable intensification strategy.
There are still few studies assessing the integrated impacts of conservation
agriculture on productivity, climate resilience, and environment/ecosystem services
conservation [39]. Long-term research and data collection and availability are
essential to technically support decision-makers for sustainably managing agricul-
tural lands [59, 117]. Moreover, there is a need for investments to encourage
pasturelands and croplands restoration while encouraging studies for a better under-
standing of the compound impacts of soil ecosystem services deterioration on
productivity and food security. The Brazilian government has taken the first steps
toward addressing this problem through the Low-Carbon Agriculture Program,
which establishes a line of credit for mainly pasture restoration and agroforestry
systems implementation.

6 Salinity

Degraded soils by salinization and sodification processes in Brazil are grossly


estimated to cover 16 Mha or 2% of the country’s land [118]. The occurrence is
restricted to the low lands surrounding the lakes of the State of Rio Grande do Sul
(RS), some areas in the Pantanal, and predominantly in the semi-arid northeastern
region. Saline, sodic, and saline-sodic soils have reduced productivity. It is very
difficult to prevent and even more difficult to recover the soil after salinization [119].
The typical semi-arid scarce rainfall and high evapotranspiration rate associated
with poor soil drainage are responsible for primary salinity in the northeast. The
salinization is enhanced by expanding irrigated crops in coastal areas where saline
water is used for irrigation, being a major threat to the soil and the crop yield in the
region [120]. Due to its low cost and abundance, gypsum is the most common
amendment to correct sodification in irrigated areas in the northeast [121, 122].
112 P. T. S. Oliveira et al.

The soils in the State of Rio Grande do Sul are barely susceptible to soil
salinization due to the large amounts of rainfall that leach the soluble salts, with an
exception for the coastal plains, particularly surrounding the Patos Lagoon which
borders the Atlantic Ocean [123, 124]. There, the seawater enters the lagoon in the
driest months, which coincides with the irrigation cycles of the flooded rice crops
[123, 125, 126]. The Patos Lagoon, along with the coastal rivers, can act as a salinity
source for the soil as it is a major water source for irrigated rice crops in the region.
Some of the alternatives used to reduce the exchangeable sodium percentage in rice
crops are the application of gypsum, limestone, rice husk, manure, vinasse, and other
sulfur-based amendments [123, 124, 126].
In Nhecolândia, a sub-region of the Pantanal, the atypical freshwater entry is
transforming the soils around the saline lakes into saline-sodic, while around the
brackish lakes soils are turning into sodic and degraded sodic soils [127]. Furquim
et al. [128] point out that the region has been through a general desalinization, as
salinization and solonization processes around the saline lakes have been replaced by
solodization, which generates degraded sodic soils. Besides excess exchangeable
sodium, degraded sodic soils also contain exchangeable hydrogen and therefore
present an acidic reaction. The authors point out that this replacement process is
strictly related to water geochemistry.
The identification of the areas susceptible to salinization/sodification is vital for
preventing enduring degradation, in order to apply proper soil and water manage-
ment. For instance, Carmona [123] has mapped the saline soils in the coastal
lowlands of Southern Brazil by field sampling. The monitoring is pertinent to the
rice production in the region. Researchers have been trying to develop remote
sensing strategies to monitor salinity soil areas [129–131]. However, studies on
salinization/sodification are scarce for Brazilian soils and require further spatial and
characterization investigations, in order to classify proper areas for irrigation and for
the reclamation of degraded soils.

7 Agrochemicals Use

In Brazil, the demand for the use of agrochemicals is mainly due to the different
crops and the diverse climatic conditions [132] that favor the emergence of living
beings that are considered harmful to agricultural production [133, 134]. In the
Brazilian territory, the most diverse types of cultivation are distributed, such as
grains, vegetables, and livestock [135].
The agrochemicals use in Brazil began in the 1950s when an agricultural mod-
ernization movement emerged [136]. The favorable context for the use of agro-
chemicals was part of a series of agricultural technologies, such as the use of
synthetic fertilizers, limestone, tractors, certified seeds, and other agricultural imple-
ments [136]. In addition, the Brazilian government started to encourage farmers,
creating technical assistance offices and offering rural credit [137]. In the 1980s,
Agricultural Land Degradation in Brazil 113

criticism of agrochemicals began to emerge, with the consequent emergence and


defense of an alternative, organic, or agroecological agriculture [136, 138].
The increase in the consumption of agrochemicals is directly related to the
continuous growth of the world population, which demands greater agricultural
production. This increased production will make use of more agrochemicals, inten-
sive irrigation, and plant varieties with better yields [138]. Currently, reports point to
Brazil as one of the largest consumers of agrochemicals in the world
[139, 140]. These studies take into account absolute numbers, number per cultivated
area, and volume of agricultural production, which, from the point of view of
industries, are wrong comparisons [141].
Among the various types of agrochemicals used, there are bactericides with the
function of destroying, eradicating, or preventing the appearance of bacteria; fungi-
cides, which serve to control, destroy, and regulate the effects of fungi; herbicides,
used to destroy weeds or any type of unwanted vegetation; and insecticides, used to
destroy, eradicate, or stun, inhibit feeding, or prevent infections and insect attack
[142]. Despite being important for increasing food production, the use of agrochem-
icals is one of the most important and difficult problems of soil contamination, which
can generate serious health problems; among the main ones are alterations in renal
functions [143, 144].
Some points must be taken into account when indicating Brazil as one of the
largest consumers of agrochemicals in the world. Brazil is one of the main agricul-
tural exporters in the world, the largest grain producer, and the largest exporter of
beef, so the tendency to produce a lot is to consume more agrochemicals [141]. In
addition, due to the extensive cultivated area and the tropical climate, agricultural
production can range from two to three crops per year [145].
In general, when used according to the recommended dose, the use of agrochem-
icals acts to control pests and diseases that damage crops, ensuring increased
productivity, reducing product prices [146, 147]. On the other hand, the use of
agrochemicals is associated with several problems, such as the residues left in bee
products, which affect agricultural pollination, groundwater contamination from
residues of pesticides, cases of poisoning, and the appearance of heavy metals in
crops, among others [148–151].
The control of agrochemical use is regulated by Law N°. 7.802/1989 [152], which
provides for research, production, use, final destination of agrochemical residues,
and packaging, among others. Parameters are also inspected by the National Health
Surveillance Agency (Agência Nacional de Vigilância Sanitária – ANVISA), the
Ministry of Agriculture, Livestock and Supply (Ministério da Agricultura, Pecuária e
Abastecimento – MAPA), and the Brazilian Institute for the Environment and
Renewable Natural Resources (Instituto Brasileiro do Meio Ambiente e dos
Recursos Naturais Renováveis – IBAMA).
114 P. T. S. Oliveira et al.

8 Microplastics

The release of plastic into the environment is known as a major threat to marine
biota, besides freshwater and estuarine environments, but the effects of microplastics
on terrestrial ecosystems remain largely unexplored [153]. Studies show that with
several potential exposure pathways, soils may represent the largest global environ-
mental reservoirs of microplastics, as plastic release to land rate is estimated at 4–23
times that released to oceans [154, 155].
Brazil ranks fourth in the production of plastic waste in the world with ~11
million tons of waste per year and, of this total, approximately 91% is collected, but
only 1.4% is actually recycled [156], threatening the soil and water ecosystems as
mismanaged waste is the principal source of plastic to the environment
[157]. According to He et al. [158], agricultural inputs from practices such as the
utilization of sewage sludge treatment and plastic mulching, besides the usage of
untreated wastewater for irrigation, represent the main source of microplastics to
soils.
In Brazil, agricultural use of sewage sludge is restricted to the states of São Paulo,
Paraná, Rio Grande do Sul, Distrito Federal, and Espírito Santo [159]. Basic sani-
tation in Brazil is precarious: almost half of the municipalities do not have sewage
treatment [160]. Even though the agricultural use of sludge portrays a sustainable
solution for sludge disposal, the sewage treatment process does not guarantee the
complete removal of microplastic particles, being a potential pathway to agricultural
soils.
The benefits of pest control, moisture retention, increased productivity, and
protection from sun and wind make plastic mulching a common and emerging
practice in vegetable and fruit cultivation in Brazil [161, 162]. However, little is
known about the plastic mulching effects in Brazil regarding microplastics. Recent
studies reinforce the need for truly biodegradable plastic mulching [163, 164].
Studies on microplastics in Brazil began in the 1970s, but the first publication was
four decades later by Ivar et al. [165] reporting the occurrence of microplastics on the
beaches of the Fernando de Noronha Archipelago. Since then, several investigations
have been published in different matrices, predominantly aquatic: microplastic
contamination on beaches, estuaries, mangroves, lotic environments, and
organisms [157].
Pollution by microplastics is an emerging topic in global soil science and a
knowledge gap to be filled. The scarcity of data on soil pollution by microplastics
for a food exporter country as Brazil makes urgent and clear the necessity of further
studies on identification, quantification, qualification, and removal techniques of
microplastics in terrestrial environments, with a focus on agricultural soils.
Agricultural Land Degradation in Brazil 115

9 Challenges and Future Perspectives for Agriculture


and Land Degradation in Brazil

In the last 50 years, Brazil has turned the tables in food production by overcoming its
own environmental, social, and economic adversities. Brazil’s successful transition
from a net food importer to a leading global food exporter was underpinned by
agricultural technology advances. The improvements were based on more conser-
vative land management policies, substituting the practice of clearing forest areas
with the recovery of degraded land for planting.
In this context, conservation practices have also played a fundamental role in the
Brazilian food production by adopting practices such as no-tillage, terracing, inte-
grated crop–livestock and crop–livestock–forestry systems, acidity correction, agri-
cultural machinery, and genetic improvement of crops for adaption to the country’s
pedoclimatic conditions [3]. Moreover, slash-and-burn practice is in decline in
Brazil [93, 95]; it is still adopted mainly by smallholders due to the lack of
inspection, besides cultural and financial reasons [92, 95]. Funding and incentives
should be provided to local farmers to combat this practice, as well as environmental
education.
Although Brazil has managed to achieve a prominent position in the global food
market, the country is still subject to a series of threats to its agricultural soils. Soil
erosion is a major threat as it removes the fertile topsoil, reduces soil productivity,
and creates a series of off-site issues by the transportation and deposition of
agricultural pollutants into water bodies. Soil erosion studies are fundamental to
the sustainable management of soil resources. Measured soil loss data are necessary
for erosion models’ development and validation. Yet, the number of studies with
observed data has decreased in Brazil [59]. Research and governmental institutions
should join forces to keep experimental plots in activity to develop a robust soil
erosion database. On the other hand, the number of studies with simulated or
estimated data is increasing, providing a better perspective of the status of soil loss
impacts in large-scale studies [61]. While water erosion studies are growing, wind
erosion comprises an important research gap in Brazil.
With the growth of the world population and increased food demand, Brazil’s
soils are inevitably necessary to guarantee food security. Aiming at increasing
agricultural production, Brazilian agriculture has bet on intensification practices to
reduce losses caused by soil degradation. Nonetheless, agricultural intensification
has contributed mainly to pasture degradation due to overgrazing. Nearly one-fifth of
the country’s territory comprises pasturelands, of which about 50% is somewhat
degraded [13]. Studies on best management practices for a sustainable intensification
are increasing in Brazil [10, 14, 104, 166]; however, more research is needed in order
to investigate particular characteristics of Brazil – which has continental proportions
and, consequently, diverse pedological and edaphoclimatic features – for an ade-
quate and adaptive agricultural management. Furthermore, ecosystem services other
than food provision need to be encouraged in research related to sustainability in
agriculture in Brazil. A better understanding of synergies and trade-offs between
116 P. T. S. Oliveira et al.

agricultural activities and the surrounding ecosystem services (e.g., water and air
quality, climate regulation, and biodiversity) is required to reduce the footprints of
agriculture in Brazil and worldwide.
To keep its soil productivity, Brazil relies on the use of agrochemicals such as
herbicides, pesticides, and fertilizers. In fact, Brazil is placed as the world’s biggest
importer of pesticides [167], which raises concern, since pesticides may be a major
environmental risk [134, 168, 169]. While the adverse effects of agrochemicals on
the health of humans and the environment are still nebulous and controversial, the
usage of agrochemicals must strictly follow regulations rules, especially because
fertilizer overuse also causes land degradation [109]. Continuous monitoring and
further investigations on behavior and fate of agrochemicals in soil are needed to
bridge this research gap.
Heavy metals accumulation in agricultural soils harms soil and water quality, in
addition to human and animal health and food security, agricultural inputs being a
major source for these contaminants [170–173]. Some studies have focused on
setting baseline concentrations [174–176] and on remediation, extraction, and indi-
cators [177–180], but we did not find studies aiming at the current status of
contamination of Brazilian agricultural soils. Once Brazil is a leading agrochemicals
consumer, contamination by heavy metals in the country comprises an important
research theme that has received little attention. In this sense, establishing soil
sampling sites for a heavy metals monitoring program is recommended.
While several studies have addressed contamination by microplastics in aquatic
environments in Brazil [157], little is known about the effects of these pollutants on
the soil. Considering that Brazil is one of the world’s top-four producers of plastic
waste [156], that mismanaged waste is the principal source of plastic to the envi-
ronment [157], and that soils may be the largest global environmental reservoirs of
microplastics [154, 155], studies focusing on soil pollution by microplastics repre-
sents an important research gap in Brazil.
Another important research gap that threatens agricultural soils comprises salini-
zation and sodification. These processes relate to poor soil drainage, high evapo-
transpiration, and scarce rainfall and can also be human induced by associating
irrigation and inadequate drainage. It is very difficult to recover the soil after
salinization [119], so assuming a preventive behavior is decisive. The studies on
salt-affected soils in Brazil are scarce. Remote sensing strategies to monitor salt-
affected soil areas have been developed [129–131], but require further spatial and
characterization investigations, in order to classify proper areas for irrigation and for
the reclamation of degraded soils.
Further studies on agricultural land productivity and degradation should be
conducted in the context of climate changes and water–food–energy security
nexus [1, 181–183]. Although Brazil has abundant water resources, the changing
climate brings important variations in temperature and rainfall patterns with
increased frequency of drought that, coupled with the increasing demand for agri-
cultural water, enhance water stress scenarios [184, 185]. With the required land
availability and agricultural technology, Brazil has the opportunity to maintain its
prominence in the agricultural goods market [13]. However, the health of its
Agricultural Land Degradation in Brazil 117

agricultural soils is a critical limiting factor. Moreover, Brazil needs to keep up to


date with the constantly evolving technology scenario. With the emergency of smart
agriculture technology, Brazil has the opportunity to make its livestock and crop
production more efficient. For that, further investments in education and smart
farming technology development are of paramount importance for overcoming the
limitations of systems integration and rural workers’ technological ability.

Acknowledgments This study was supported by grants from the Ministry of Science, Technology,
Innovation and Communication (MCTIC) and the National Council for Scientific and Technolog-
ical Development (CNPq) [grant numbers 441289/2017-7 and 309752/2020-5]. This study was also
financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior – Brazil
(CAPES) – Finance Code 001 and CAPES PrInt.

References

1. Oliveira PTS, Almagro A, Colman CB, Kobayashi ANA, Rodrigues DBB, Meira Neto AA,
Gupta HV (2019) Nexus of water-food-energy-ecosystem services in the Brazilian
Cerrado. In: Vieira da Silva RC, CEM T, Scott CA (eds) Water and climate – modeling in
large basins. ABRHidro, Porto Alegre, pp 7–30
2. CEPEA (2021) PIB-AGRO/CEPEA: COM AVANÇO DE 24,3% NO ANO, PIB AGRO
ALCANÇA PARTICIPAÇÃO DE 26,6% NO PIB BRASILEIRO EM 2020. Centro de
Estudos Avançados em Economia Aplicada. https://cepea.esalq.usp.br/br/releases/pib-agro-
cepea-com-avanco-de-24-3-no-ano-pib-agro-alcanca-participacao-de-26-6-no-pib-brasileiro-
em-2020.aspx. Accessed 6 Jan 2021
3. Embrapa (2018) Visão 2030: o futuro da agricultura brasileira. Brasília, 212 p. https://www.
Embrapa.br/visao/o-futuro-da-agricultura-brasileira. Accessed 6 Jan 2021
4. Bertol I (2016) Conservação do solo no Brasil: histórico, situação atual e o que esperar para o
futuro. Soc Bras Ciênc Solo. https://www.sbcs.org.br/?noticia_geral=a-conservacao-do-solo-
no-brasil. Accessed 6 Jan 2021
5. Alexandratos N, Bruinsma J (2012) World agriculture towards 2030/2050: the 2012 revision.
ESA working paper no. 12-03. FAO, Rome
6. Pastor AV, Palazzo A, Havlik P, Biemans H, Wada Y, Obersteiner M, Kabat P, Ludwig F
(2019) The global nexus of food–trade–water sustaining environmental flows by 2050. Nat
Sustain 2:499–507. https://doi.org/10.1038/s41893-019-0287-1
7. Hecht SB (1993) The logic of livestock and deforestation in Amazonia. Bioscience 43(10):
687–695
8. Gibbs HK, Ruesch AS, Achard F, Clayton MK, Holmgren P, Ramankutty N, Foley JA (2010)
Tropical forests were the primary sources of new agricultural land in the 1980s and 1990s.
Proc Natl Acad Sci 107(38):16732–16737
9. Ramalho-Filho A, Freitas PL, Claessen MEC (2009) Land degradation and the zero-tillage
system in Brazil. In: Cano AF, Mermut AR, Arocena JM, Silla RO (eds) Land degradation and
rehabilitation – dryland ecosystems. Schweizerbart Science Publishers, pp 311–324
10. Sone JS, Oliveira PTS, Zamboni PAP, Motta Vieira NOM, Carvalho GA, Macedo MCM,
Araujo AR, Montagner DB, Alves Sobrinho T (2019) Effects of long-term crop-livestock-
forestry systems on soil erosion and water infiltration in a Brazilian Cerrado site. Sustainability
11(19):5339. https://doi.org/10.3390/su11195339
11. MapBiomas (2021) Projeto MapBiomas – Coleção 5 da Série Anual de Mapas de Cobertura e
Uso de Solo do Brasil. https://mapbiomas.org/. Accessed Mar 2021
118 P. T. S. Oliveira et al.

12. FAOSTAT (2021) Food and agriculture data. Food and Agriculture Organization of the United
Nations. http://www.fao.org/faostat/. Accessed Mar 2021
13. Strassburg BB, Latawiec AE, Barioni LG, Nobre CA, Da Silva VP, Valentim JF et al (2014)
When enough should be enough: improving the use of current agricultural lands could meet
production demands and spare natural habitats in Brazil. Glob Environ Chang 28:84–97.
https://doi.org/10.1016/j.gloenvcha.2014.06.001
14. Latawiec AE, Strassburg BB, Valentim JF, Ramos F, Alves-Pinto HN (2014) Intensification of
cattle ranching production systems: socioeconomic and environmental synergies and risks in
Brazil. Animal 8(8):1255–1263. https://doi.org/10.1017/S1751731114001566
15. Martinelli LA, Naylor R, Vitousek PM, Moutinho P (2010) Agriculture in Brazil: impacts,
costs, and opportunities for a sustainable future. Curr Opin Environ Sustain 2(5–6):431–438.
https://doi.org/10.1016/j.cosust.2010.09.008
16. Godfray HCJ, Beddington JR, Crute IR, Haddad L, Lawrence D, Muir JF et al (2010) Food
security: the challenge of feeding 9 billion people. Science 327(5967):812–818. https://doi.
org/10.1126/science.1185383
17. Govers G, Merckx R, Wesemael BV, Oost KV (2017) Soil conservation in the 21st century:
why we need smart agricultural intensification. Soil 3(1):45–59. https://doi.org/10.5194/soil-
3-45-2017
18. Alves BJ, Madari BE, Boddey RM (2017) Integrated crop–livestock–forestry systems: pros-
pects for a sustainable agricultural intensification. Nutr Cycl Agroecosyst 108(1):1–4. https://
doi.org/10.1007/s10705-017-9851-0
19. Struik PC, Kuyper TW (2017) Sustainable intensification in agriculture: the richer shade of
green. A review. Agron Sustain Dev 37(5):1–15. https://doi.org/10.1007/s13593-017-0445-7
20. Conab (2021) Harvest historical data. The Brazilian National Company of Food Supply.
https://www.conab.gov.br/. Accessed Mar 2021
21. Rada N (2013) Assessing Brazil’s Cerrado agricultural miracle. Food Policy 38:146–155.
https://doi.org/10.1016/j.foodpol.2012.11.002
22. Gomes L, Simões SJ, Dalla Nora EL, de Sousa-Neto ER, Forti MC, Ometto JPH (2019)
Agricultural expansion in the Brazilian Cerrado: increased soil and nutrient losses and
decreased agricultural productivity. Land 8(1):12. https://doi.org/10.3390/land8010012
23. Zalles V, Hansen MC, Potapov PV, Stehman SV, Tyukavina A, Pickens A et al (2019) Near
doubling of Brazil’s intensive row crop area since 2000. Proc Natl Acad Sci 116(2):428–435.
https://doi.org/10.1073/pnas.1810301115
24. Gibbs HK, Rausch L, Munger J, Schelly I, Morton DC, Noojipady P et al (2015) Brazil’s soy
moratorium. Science 347(6220):377–378. https://doi.org/10.1126/science.aaa0181
25. Oliveira PTS, Nearing MA, Wendland E (2015) Orders of magnitude increase in soil erosion
associated with land use change from native to cultivated vegetation in a Brazilian savannah
environment. Earth Surf Process Landf 40(11):1524–1532. https://doi.org/10.1002/esp.3738
26. Grecchi RC, Gwyn QHJ, Bénié GB, Formaggio AR, Fahl FC (2014) Land use and land cover
changes in the Brazilian Cerrado: a multidisciplinary approach to assess the impacts of
agricultural expansion. Appl Geogr 55:300–312. https://doi.org/10.1016/j.apgeog.2014.
09.014
27. Filoso S, do Carmo JB, Mardegan SF, Lins SRM, Gomes TF, Martinelli LA (2015)
Reassessing the environmental impacts of sugarcane ethanol production in Brazil to help
meet sustainability goals. Renew Sust Energ Rev 52:1847–1856. https://doi.org/10.1016/j.
rser.2015.08.012
28. Martinelli LA, Filoso S (2008) Expansion of sugarcane ethanol production in Brazil: environ-
mental and social challenges. Ecol Appl 18(4):885–898. https://doi.org/10.1890/07-1813.1
29. Nehren UDO, Kirchner A, Sattler D, Turetta AP, Heinrich J (2013) Impact of natural climate
change and historical land use on landscape development in the Atlantic Forest of Rio de
Janeiro, Brazil. An Acad Bras Cienc 85(2):497–518
Agricultural Land Degradation in Brazil 119

30. Bordonal RO, Carvalho JLN, Lal R, de Figueiredo EB, de Oliveira BG, La Scala N (2018)
Sustainability of sugarcane production in Brazil. A review. Agron Sustain Dev 38(2):1–23.
https://doi.org/10.1007/s13593-018-0490-x
31. Teixeira HM, Cardoso IM, Bianchi FJ, da Cruz Silva A, Jamme D, Peña-Claros M (2020)
Linking vegetation and soil functions during secondary forest succession in the Atlantic forest.
For Ecol Manag 457:117696. https://doi.org/10.1016/j.foreco.2019.117696
32. Rodrigues RR, Lima RAF, Gandolfi S, Nave AG (2009) On the restoration of high diversity
forests: 30 years of experience in the Brazilian Atlantic Forest. Biol Conserv 142(6):
1242–1251. https://doi.org/10.1016/j.biocon.2008.12.008
33. Joly CA, Rodrigues RR, Metzger JP, Haddad CFB, Verdade LM, Oliveira MC, Bolzani VS
(2010) Biodiversity conservation research, training, and policy in Sao Paulo. Science
328(5984):1358–1359. https://doi.org/10.1126/science.1188639
34. Mendes MS, Latawiec AE, Sansevero JB, Crouzeilles R, Moraes LF, Castro A et al (2019)
Look down – here is a gap – the need to include soil data in Atlantic Forest restoration. Restor
Ecol 27(2):361–370. https://doi.org/10.1111/rec.12875
35. Morton DC, DeFries RS, Shimabukuro YE, Anderson LO, Arai E, del Bon Espirito-Santo F
et al (2006) Cropland expansion changes deforestation dynamics in the southern Brazilian
Amazon. Proc Natl Acad Sci 103(39):14637–14641. https://doi.org/10.1073/pnas.
0606377103
36. Barona E, Ramankutty N, Hyman G, Coomes OT (2010) The role of pasture and soybean in
deforestation of the Brazilian Amazon. Environ Res Lett 5(2):024002. https://doi.org/10.1088/
1748-9326/5/2/024002
37. Eri M, da Silva Junior CA, Lima M, Júnior NLS, de Oliveira-Júnior JF, Teodoro PE et al
(2020) Capitalizing on opportunities provided by pasture sudden death to enhance livestock
sustainable management in Brazilian Amazonia. Environ Dev 33:100499. https://doi.org/10.
1016/j.envdev.2020.100499
38. Silva MA, Lima M, Silva Junior CA, Costa GM, Peres CA (2018) Achieving low-carbon cattle
ranching in the Amazon: ‘pasture sudden death’ as a window of opportunity. Land Degrad
Dev 29(10):3535–3543. https://doi.org/10.1002/ldr.3087
39. Merten GH, Minella JP (2013) The expansion of Brazilian agriculture: soil erosion scenarios.
Int Soil Water Conserv Res 1(3):37–48. https://doi.org/10.1016/S2095-6339(15)30029-0
40. Terraclass (2016) Avaliação da dinâmica do uso e cobertura da terra no período de 10 anos nas
áreas desflorestadas da Amazônia legal Brasileira. https://ainfo.cnptia.Embrapa.br/digital/
bitstream/item/152807/1/TerraClass.pdf
41. Fearnside PM (2005) Deforestation in Brazilian Amazonia: history, rates, and consequences.
Conserv Biol 19(3):680–688. https://doi.org/10.1111/j.1523-1739.2005.00697.x
42. Shakesby RA, Doerr SH (2006) Wildfire as a hydrological and geomorphological agent. Earth
Sci Rev 74(3–4):269–307. https://doi.org/10.1016/j.earscirev.2005.10.006
43. Flores BM, Holmgren M, Xu C, Van Nes EH, Jakovac CC, Mesquita RC, Scheffer M (2017)
Floodplains as an Achilles’ heel of Amazonian forest resilience. Proc Natl Acad Sci 114(17):
4442–4446. https://doi.org/10.1073/pnas.1617988114
44. Godoi RF, Rodrigues DB, Borrelli P, Oliveira PTS (2021) High-resolution soil erodibility map
of Brazil. Sci Total Environ 781:146673. https://doi.org/10.1016/j.scitotenv.2021.146673
45. Oliveira PTS, Wendland E, Nearing MA (2013) Rainfall erosivity in Brazil: a review. Catena
100:139–147. https://doi.org/10.1016/j.catena.2012.08.006
46. Almagro A, Oliveira PTS, Nearing MA, Hagemann S (2017) Projected climate change impacts
in rainfall erosivity over Brazil. Sci Rep 7(1):1–12. https://doi.org/10.1038/s41598-017-
08298-y
47. Flores BM, Staal A, Jakovac CC, Hirota M, Holmgren M, Oliveira RS (2020) Soil erosion as a
resilience drain in disturbed tropical forests. Plant Soil 450:11–25. https://doi.org/10.1007/
s11104-019-04097-8
48. Overbeck GE, Vélez-Martin E, Scarano FR, Lewinsohn TM, Fonseca CR, Meyer ST, Müller
SC, Ceotto P, Dadalt L, Durigan G, Ganade G, Gossner MM, Guadagnin DL, Lorenzen K,
120 P. T. S. Oliveira et al.

Jacobi CM, Weisser WW, Pillar VD (2015) Conservation in Brazil needs to include non-forest
ecosystems. Divers Distrib 21:1455–1460. https://doi.org/10.1111/ddi.12380
49. Seidl AF, de Silva JDSV, Moraes AS (2001) Cattle ranching and deforestation in the Brazilian
Pantanal. Ecol Econ 36(3):413–425. https://doi.org/10.1016/S0921-8009(00)00238-X
50. Overbeck GE, Müller SC, Fidelis A, Pfadenhauer J, Pillar VD, Blanco CC et al (2007) Brazil’s
neglected biome: the South Brazilian Campos. Perspect Plant Ecol Evol Syst 9(2):101–116.
https://doi.org/10.1016/j.ppees.2007.07.005
51. Nabinger C, Ferreira ET, Freitas AK, Carvalho PDF, Sant’Anna DM (2009) Produção animal
com base no campo nativo: aplicações de resultados de pesquisa. In: Campos Sulinos:
conservação e uso sustentável da biodiversidade, pp 175–198
52. Oliveira TE, Freitas DS, Gianezini M, Ruviaro CF, Zago D, Mércio TZ et al (2017) Agricul-
tural land use change in the Brazilian Pampa biome: the reduction of natural grasslands. Land
Use Policy 63:394–400. https://doi.org/10.1016/j.landusepol.2017.02.010
53. IBGE (2017) Censo Agropecuário 2017. Brazilian Institute of Geography and Statistics.
https://censoagro2017.ibge.gov.br/. Accessed Mar 2021
54. Roesch LFW, Vieira FCB, Pereira VA, Schünemann AL, Teixeira IF, Senna AJT, Stefenon
VM (2009) The Brazilian Pampa: a fragile biome. Diversity 1(2):182–198. https://doi.org/10.
3390/d1020182
55. Junk WJ, Da Cunha CN (2012) Pasture clearing from invasive woody plants in the Pantanal: a
tool for sustainable management or environmental destruction? Wetl Ecol Manag 20(2):
111–122. https://doi.org/10.1007/s11273-011-9246-y
56. Lapola DM, Martinelli LA, Peres CA, Ometto JP, Ferreira ME, Nobre CA et al (2014)
Pervasive transition of the Brazilian land-use system. Nat Clim Chang 4(1):27–35. https://
doi.org/10.1038/nclimate2056
57. Colman CB, Oliveira PTS, Almagro A, Soares-Filho BS, Rodrigues DB (2019) Effects of
climate and land-cover changes on soil erosion in Brazilian Pantanal. Sustainability 11(24):
7053. https://doi.org/10.3390/su11247053
58. Lucas MC, Kublik N, DBB R, Meira Neto AA, Almagro A, DdCD M, Zipper SC, PTS O
(2021) Significant baseflow reduction in the Sao Francisco River basin. Water 13(1):2. https://
doi.org/10.3390/w13010002
59. Anache JA, Wendland EC, Oliveira PT, Flanagan DC, Nearing MA (2017) Runoff and soil
erosion plot-scale studies under natural rainfall: a meta-analysis of the Brazilian experience.
Catena 152:29–39. https://doi.org/10.1016/j.catena.2017.01.003
60. Wischmeier WH, Smith DD (1978) Predicting rainfall erosion losses: a guide to conservation
planning (no. 537). Department of Agriculture, Science and Education Administration
61. Borrelli P, Lugato E, Montanarella L, Panagos P (2017) A new assessment of soil loss due to
wind erosion in European agricultural soils using a quantitative spatially distributed modelling
approach. Land Degrad Dev 28(1):335–344. https://doi.org/10.1002/ldr.2588
62. Zhang H, Yang Q, Li R, Liu Q, Moore D, He P et al (2013) Extension of a GIS procedure for
calculating the RUSLE equation LS factor. Comput Geosci 52:177–188. https://doi.org/10.
1016/j.cageo.2012.09.027
63. Zhang H, Wei J, Yang Q, Baartman JE, Gai L, Yang X et al (2017) An improved method for
calculating slope length (λ) and the LS parameters of the Revised Universal Soil Loss Equation
for large watersheds. Geoderma 308:36–45. https://doi.org/10.1016/j.geoderma.2017.08.006
64. Yamazaki D, Ikeshima D, Tawatari R, Yamaguchi T, O'Loughlin F, Neal JC et al (2017) A
high-accuracy map of global terrain elevations. Geophys Res Lett 44(11):5844–5853. https://
doi.org/10.1002/2017GL072874
65. Alewell C, Borrelli P, Meusburger K, Panagos P (2019) Using the USLE: chances, challenges
and limitations of soil erosion modelling. Int Soil Water Conserv Res 7(3):203–225. https://
doi.org/10.1016/j.iswcr.2019.05.004
66. Durigon VL, Carvalho DF, Antunes MAH, Oliveira PTS, Fernandes MM (2014) NDVI time
series for monitoring RUSLE cover management factor in a tropical watershed. Int J Remote
Sens 35(2):441–453. https://doi.org/10.1080/01431161.2013.871081
Agricultural Land Degradation in Brazil 121

67. Almagro A, Thomé TC, Colman CB, Pereira RB, Junior JM, Rodrigues DBB, Oliveira PTS
(2019) Improving cover and management factor (C-factor) estimation using remote sensing
approaches for tropical regions. Int Soil Water Conserv Res 7(4):325–334. https://doi.org/10.
1016/j.iswcr.2019.08.005
68. Macedo PMS, Oliveira PTS, Antunes MAH, Durigon VL, Fidalgo ECC, de Carvalho DF
(2021) New approach for obtaining the C-factor of RUSLE considering the seasonal effect of
rainfalls on vegetation cover. Int Soil Water Conserv Res 2021(9):207–216. https://doi.org/10.
1016/j.iswcr.2020.12.001
69. Colman CB (2018) Impacts of climate and land use changes on soil erosion in the Upper
Paraguay Basin. Master’s dissertation, Federal University of Mato Grosso do Sul, Campo
Grande, Brazil
70. Freitas PL, Landers JN (2014) The transformation of agriculture in Brazil through develop-
ment and adoption of zero tillage conservation agriculture. Int Soil Water Conserv Res 2(1):
35–46. https://doi.org/10.1016/S2095-6339(15)30012-5
71. Hobbs PR (2007) Conservation agriculture: what is it and why is it important for future
sustainable food production? J Agric Sci (Camb) 145(2):127. https://doi.org/10.1017/
S0021859607006892
72. Fuentes-Llanillo R, Telles TS, Soares Junior D, de Melo TR, Friedrich T, Kassam A (2021)
Expansion of no-tillage practice in conservation agriculture in Brazil. Soil Tillage Res 208:
104877. https://doi.org/10.1016/j.still.2020.104877
73. Embrapa (2014) Simplificação do Plantio Direto reduz eficiência da lavoura. Embrapa Solos.
https://www.Embrapa.br/busca-de-noticias/-/noticia/1909275/simplificacao-do-plantio-
direto-reduz-eficiencia-da-lavoura
74. Castro Filho CD, Henklain JC, Vieira MJ, Casão Jr R (1991) Tillage methods and soil and
water conservation in southern Brazil. Soil Tillage Res 20(2–4):271–283. https://doi.org/10.
1016/0167-1987(91)90043-W
75. ILPF (2016) ILPF em números. Embrapa, Sinop, 12 p. http://www.ilpf.com.br/
76. Salton JC, Mercante FM, Tomazi M, Zanatta JA, Concenço G, Silva WM, Retore M (2014)
Integrated crop-livestock system in tropical Brazil: toward a sustainable production system.
Agric Ecosyst Environ 190:70–79. https://doi.org/10.1016/j.agee.2013.09.023
77. Kluthcouski J, Cordeiro LAM (2018) Do Plantio Direto aos Sistemas de Integração entre
Lavoura e Pecuária: Trajetórias da Produtividade Agropecuária. In: Olhares para 2030:
Desenvolvimento Sustentável. Embrapa. https://www.Embrapa.br/olhares-para-2030
78. Fryrear DW, Bilbro JD, Saleh A, Schomberg H, Stout JE, Zobeck TM (2000) RWEQ:
improved wind erosion technology. J Soil Water Conserv 55(2):183–189
79. de Jong P, Tanajura CAS, Sánchez AS, Dargaville R, Kiperstok A, Torres EA (2018)
Hydroelectric production from Brazil’s São Francisco River could cease due to climate change
and inter-annual variability. Sci Total Environ 634:1540–1553. https://doi.org/10.1016/j.
scitotenv.2018.03.256
80. de Jong P, Barreto TB, Tanajura CAS, Kouloukoui D, Oliveira-Esquerre KP, Kiperstok A,
Torres EA (2019) Estimating the impact of climate change on wind and solar energy in Brazil
using a South American regional climate model. Renew Energy 141:390–401. https://doi.org/
10.1016/j.renene.2019.03.086
81. Fenta AA, Tsunekawa A, Haregeweyn N, Poesen J, Tsubo M, Borrelli P, Panagos P,
Vanmaercke M, Broeckx J, Yasuda H, Kawai T, Kurosaki Y (2020) Land susceptibility to
water and wind erosion risks in the East Africa region. Sci Total Environ 703:135016. https://
doi.org/10.1016/j.scitotenv.2019.135016
82. Borrelli P, Alewell C, Alvarez P, Anache JAA, Baartman J, Ballabio C et al (2021) Soil
erosion modelling: a global review and statistical analysis. Sci Total Environ 780:146494.
https://doi.org/10.1016/j.scitotenv.2021.146494
83. Lisboa B, Lattuada D (2019) ARTIGO: Dia Mundial do Solo alerta para os perigos da erosão.
Secretaria da Agricultura, Pecuária e Desenvolvimento Rural. https://www.agricultura.rs.gov.
br/artigo-dia-mundial-do-solo-alerta-para-os-perigos-da-erosao
122 P. T. S. Oliveira et al.

84. Rovedder APM, Eltz FLF (2008) Revegetation with cover crops for soils under arenization
and wind erosion in Rio Grande do Sul state, Brazil. Rev Bras Ciênc Solo 32(1):315–321.
https://doi.org/10.1590/S0100-06832008000100029
85. Buschiazzo DE, Zobeck TM (2008) Validation of WEQ, RWEQ and WEPS wind erosion for
different arable land management systems in the Argentinean pampas. Earth Surf Process
Landf 33(12):1839–1850. https://doi.org/10.1002/esp.1738
86. Guo Z, Zobeck TM, Zhang K, Li F (2013) Estimating potential wind erosion of agricultural
lands in northern China using the Revised Wind Erosion Equation and geographic information
systems. J Soil Water Conserv 68(1):13–21. https://doi.org/10.2489/jswc.68.1.13
87. Instituto Nacional de Meteorologia – INMET. portal.inmet.gov.br
88. Almagro A, Oliveira PTS, Meira Neto AA, Roy T, Troch P (2021) CABra: a novel large-
sample dataset for Brazilian catchments. Hydrol Earth Syst Sci 25(6):3105–3135. https://doi.
org/10.5194/hess-25-3105-2021
89. Goossens D (2003) The on-site and off-site effects of wind erosion. In: Wind erosion on
agricultural land in Europe, pp 29–38
90. Adams C, Chamlian Munari L, Van Vliet N, Sereni Murrieta RS, Piperata BA, Futemma C,
Novaes Pedroso N, Santos Taqueda C, Abrahão Crevelaro M, Spressola-Prado VL (2013)
Diversifying incomes and losing landscape complexity in Quilombola shifting cultivation
communities of the Atlantic rainforest (Brazil). Hum Ecol 41:119–137. https://doi.org/10.
1007/s10745-012-9529-9
91. Peroni N, Hanazaki N (2002) Current and lost diversity of cultivated varieties, especially
cassava, under swidden cultivation systems in the Brazilian Atlantic Forest. Agric Ecosyst
Environ 92:171–183. https://doi.org/10.1016/S0167-8809(01)00298-5
92. Serrão EAS, Nepstad D, Walker R (1996) Upland agricultural and forestry development in the
Amazon: sustainability, criticality and resilience. Ecol Econ 18:3–13. https://doi.org/10.1016/
0921-8009(95)00092-5
93. van Vliet N, Adams C, Vieira ICG, Mertz O (2013) “Slash and burn” and “shifting” cultivation
systems in forest agriculture frontiers from the Brazilian Amazon. Soc Nat Resour 26:1454–
1467. https://doi.org/10.1080/08941920.2013.820813
94. Galford GL, Soares-Filho B, Cerri CEP (2013) Prospects for land-use sustainability on the
agricultural frontier of the Brazilian Amazon. Philos Trans R Soc B Biol Sci 368. https://doi.
org/10.1098/rstb.2012.0171
95. Edivaldo T, Rosell S (2020) Slash-and-burn agriculture in southern Brazil: characteristics,
food production and prospects. Scott Geogr J 136:176–194. https://doi.org/10.1080/
14702541.2020.1776893
96. de Majo C, Moretto SP (2021) From slash and burn to winemaking: the historical trajectory of
Italian colonos in the uplands of Rio Grande do Sul, Brazil. Mod Italy 26:141–158. https://doi.
org/10.1017/mit.2021.23
97. Mamede MA, de Araújo FS (2008) Effects of slash and burn practices on a soil seed bank of
caatinga vegetation in Northeastern Brazil. J Arid Environ 72:458–470. https://doi.org/10.
1016/j.jaridenv.2007.07.014
98. Comte I, Davidson R, Lucotte M, de Carvalho CJR, de Assis Oliveira F, da Silva BP,
Rousseau GX (2012) Physicochemical properties of soils in the Brazilian Amazon following
fire-free land preparation and slash-and-burn practices. Agric Ecosyst Environ 156:108–115.
https://doi.org/10.1016/j.agee.2012.05.004
99. Tomich TP, Noordwijk M, Vosti SA, Witcover J (1998) Agricultural development with
rainforest conservation: methods for seeking best bet alternatives to slash-and-burn, with
applications to Brazil and Indonesia. Agric Econ 19:159–174. https://doi.org/10.1111/j.
1574-0862.1998.tb00523.x
100. Tremblay S, Lucotte M, Revéret JP, Davidson R, Mertens F, Passos CJS, Romaña CA (2015)
Agroforestry systems as a profitable alternative to slash and burn practices in small-scale
agriculture of the Brazilian Amazon. Agrofor Syst 89:193–204. https://doi.org/10.1007/
s10457-014-9753-y
Agricultural Land Degradation in Brazil 123

101. FAO and ITPS (2015) Status of the world’s soil resources (SWSR) – main report. Food and
Agriculture Organization of the United Nations and Intergovernmental Technical Panel on
Soils, Rome
102. Sartori M, Philippidis G, Ferrari E, Borrelli P, Lugato E, Montanarella L, Panagos P (2019) A
linkage between the biophysical and the economic: assessing the global market impacts of soil
erosion. Land Use Policy 86:299–312. https://doi.org/10.1016/j.landusepol.2019.05.014
103. USDA-FSA (2021) Livestock and poultry: world markets and trade report. Foreign Agricul-
tural Service – United States Department of Agriculture. https://www.fas.usda.gov/data/
livestock-and-poultry-world-markets-and-trade. Accessed 12 Oct 2021
104. Sone JS, Oliveira PTS, Euclides VPB, Montagner DB, de Araujo AR, Zamboni PAP, Vieira
NOM, Carvalho GA, Sobrinho TA (2020) Effects of nitrogen fertilisation and stocking rates
on soil erosion and water infiltration in a Brazilian Cerrado farm. Agric Ecosyst Environ 304
(August):107159. https://doi.org/10.1016/j.agee.2020.107159
105. Teague WR (2018) Forages and pastures symposium: cover crops in livestock production:
whole-system approach: managing grazing to restore soil health and farm livelihoods. J Anim
Sci 96(4):1519–1530. https://doi.org/10.1093/jas/skx060
106. Euclides VPB, Carpejani GC, Montagner DB, Nascimento Junior D, Barbosa RA, Difante GS
(2018) Maintaining post-grazing sward height of Panicum maximum (cv. Mombaça) at 50 cm
led to higher animal performance compared with post-grazing height of 30 cm. Grass Forage
Sci 73:174–182. https://doi.org/10.1111/gfs.12292
107. Junior AAB, Moraes A, Veiga M, Pelissari A, Dieckow J (2009) Crop-livestock system:
intensified use of agricultural lands. Ciência Rural 39:1925–1933
108. USDA, NRCS (2019) Conservation practice standard: nutrient management – CODE
590-CPS-1. Natural Resources Conservation Service – United States Department of
Agriculture
109. Massah J, Azadegan B (2016) Effect of chemical fertilizers on soil compaction and degrada-
tion. Agric Mechaniz Asia, Afr Latin Am 47(1):44–50
110. Embrapa (2013) Balanço Social. http://bs.sede. Embrapa.br/2012/. Accessed 22 May 2013
111. IBGE (2020) Monitoramento da cobertura e uso da terra do Brasil: 2016–2018. https://
biblioteca.ibge.gov.br/
112. Pavei DS, Panachuki E, Salton JC, Sone JS, Sobrinho TA, Valim WC, Oliveira PTS (2021)
Soil physical properties and interrill erosion in agricultural production systems after 20 years
of cultivation. Rev Bras Ciênc Solo:45. https://doi.org/10.36783/18069657rbcs20210039
113. Panachuki E, Bertol I, Sobrinho TA, Oliveira PTS, Rodrigues DBB (2011) Perdas de solo e de
água e infiltração de água em latossolo vermelho sob sistemas de manejo. Rev Bras Ciênc Solo
35:1777–1785
114. Taboada MA, Micucci FG, Cosentino DJ, Lavado RS (1998) Comparison of compaction
induced by conventional and zero tillage in two soils of the Rolling Pampa of Argentina. Soil
Tillage Res 49(1–2):57–63. https://doi.org/10.1016/S0167-1987(98)00132-9
115. Franzluebbers AJ, Sawchik J, Taboada MA (2014) Agronomic and environmental impacts of
pasture–crop rotations in temperate North and South America. Agric Ecosyst Environ 190:18–
26. https://doi.org/10.1016/j.agee.2013.09.017
116. World Bank (2012) Carbon sequestration in agricultural soils. Report no. 67395-GLB. World
Bank, Washington. 85 pp
117. Manale A, Sharpley A, DeLong C, Speidel D, Gantzer C, Peterson J, Martin R, Lindahl C,
Adusumilli N (2018) Principles and policies for soil and water conservation. J Soil Water
Conserv 73:96–99
118. Ribeiro MR, Freire FJ, Montenegro AA (2003) Solos halomórficos no Brasil: Ocorrência,
gênese, classificação, uso e manejo sustentável. In: Curi N, Marques JJ, Guilherme LRG, Lima
JM, Lopes AS, Alvarez VVH (eds) Tópicos em Ciência do Solo. Sociedade Brasileira de
Ciência do Solo, Viçosa, pp 165–208
119. FAO (2015) Regional assessment of soil changes in Latin America and the Caribbean. In:
Status of the world’s soil resources. Rome, Italy: main report, pp 362–399
124 P. T. S. Oliveira et al.

120. Ribeiro MR, Ribeiro Filho MR, Jacomine PKT (2016) Origem e classificação dos solos
afetados por sais. In: Gheyi HR, Dias NS, Lacerda CF, Gomes Filho E (eds) Manejo da
salinidade na agricultura: Estudo básico e aplicados. INCTSal, Fortaleza, p 9
121. Holanda JS, Vitti GC, Salviano AAC, Medeleos JDF, Amorim JRA (1998) Chemical property
changes in saline-sodic soil by subsoiling and amendments. Rev Bras Ciênc Solo 22:387–394.
https://doi.org/10.1590/S0100-06831998000300003
122. Melo RM, Barros MFC, Santos PM, Rolim MM (2008) Reclamation of saline-sodic soils by
application of mineral gypsum. Rev Bras Eng Agric Ambient 12:376–380. https://doi.org/10.
1590/S1415-43662008000400007
123. Carmona FC (2011) Water and soil salinity and its influence on irrigated rice. Doctoral thesis,
Universidade Federal do Rio Grande do Sul, Porto Alegre, Brazil. https://www.lume.ufrgs.br/
bitstream/handle/10183/29534/000776595.pdf
124. Denardin LGO, de Campos Carmona F, Tiecher T, Martins AP, Alves LA, Weber EJ,
Anghinoni I (2018) Salt-affected soils of the coastal plains in Rio Grande do Sul, Brazil.
Geoderma Reg 14:e00186. https://doi.org/10.1016/j.geodrs.2018.e00186
125. Windom HL, Niencheski LF, Smith Jr RG (1999) Biogeochemistry of nutrients and trace
metals in the estuarine region of the Patos Lagoon (Brazil). Estuar Coast Shelf Sci 48(1):
113–123. https://doi.org/10.1006/ecss.1998.0410
126. Fraga TI, Carmona FDC, Anghinoni I, Genro Junior SA, Marcolin E (2010) Flooded rice yield
as affected by levels of water salinity in different stages of its cycle. Rev Bras Ciênc Solo 34:
175–182. https://doi.org/10.1590/S0100-06832010000100018
127. Andrade GRP, Furquim SAC, do Nascimento TTV, Brito AC, Camargo GR, de Souza GC
(2020) Transformation of clay minerals in salt-affected soils, Pantanal wetland, Brazil.
Geoderma 371:114380. https://doi.org/10.1016/j.geoderma.2020.114380
128. Furquim SAC, Santos MA, Vidoca TT, de Almeida Balbino M, Cardoso EL (2017) Salt-
affected soils evolution and fluvial dynamics in the Pantanal wetland, Brazil. Geoderma 286:
139–152. https://doi.org/10.1016/j.geoderma.2016.10.030
129. Rocha Neto OCD, Teixeira ADS, Leão RADO, Moreira LCJ, Galvão LS (2017) Hyperspectral
remote sensing for detecting soil salinization using ProSpecTIR-vs aerial imagery and sensor
simulation. Remote Sens 9(1):42. https://doi.org/10.3390/rs9010042
130. Bouaziz M, Matschullat J, Gloaguen R (2011) Improved remote sensing detection of soil
salinity from a semi-arid climate in Northeast Brazil. Compt Rendus Geosci 343(11–12):
795–803. https://doi.org/10.1016/j.crte.2011.09.003
131. Moreira LCJ, Teixeira ADS, Galvão LS (2015) Potential of multispectral and hyperspectral
data to detect saline-exposed soils in Brazil. GISci Remote Sens 52(4):416–436. https://doi.
org/10.1080/15481603.2015.1040227
132. Alvares CA, Stape JL, Sentelhas PC, de Moraes Gonçalves JL, Sparovek G (2013) Köppen’s
climate classification map for Brazil. Meteorol Z:711–728. https://doi.org/10.1127/0941-
2948/2013/0507
133. Rasmussen JJ, Reiler EM, Carazo E, Matarrita J, Muñoz A, Cedergreen N (2016) Influence of
rice field agrochemicals on the ecological status of a tropical stream. Sci Total Environ 542:
12–21. https://doi.org/10.1016/j.scitotenv.2015.10.062
134. Sparks TC, Lorsbach BA (2017) Perspectives on the agrochemical industry and agrochemical
discovery. Pest Manag Sci 73(4):672–677. https://doi.org/10.1002/ps.4457
135. Embrapa (2021) Grandes contribuições para a agricultura brasileira – Portal Embrapa. https://
www.Embrapa.br/grandes-contribuicoes-para-a-agricultura-brasileira
136. de Carvalho MMX, Nodari ES, Nodari RO (2017) “Defensivos” ou “agrotóxicos”? História do
uso e da percepção dos agrotóxicos no estado de Santa Catarina, Brasil, 1950-2002. História,
Ciências, Saúde-Manguinhos 24(1):75–91. https://doi.org/10.1590/
s0104-59702017000100002
137. de Araujo PFC, Meyer RL (1977) Agricultural credit policy in Brazil: objectives and results.
Am J Agric Econ 59(5):957–961. https://doi.org/10.2307/1239871
Agricultural Land Degradation in Brazil 125

138. Dalcin D, de Freitas JB, Padula JD, Dewes H (2014) Organic products in Brazil: from an
ideological orientation to a market choice. Br Food J 116(12):1998–2015. https://doi.org/10.
1108/BFJ-01-2013-0008
139. Bombardi (2017) Atlas Geográfico do Uso de Agrotóxicos no Brasil e Conexões com a União
Europeia | Ecotox Brasil. https://ecotoxbrasil.org.br/comunicacao-cientifica/8/atlas-
geografico-do-uso-de-agrotoxicos-no-brasil-e-conexoes-com-a-uniao-europeia/
140. Londres F (2012) Agrotóxicos no Brasil: Um guia para ação em defesa da vida (2a edição).
Rede Brasileira de Justiça Ambiental: Articulação Nacional de Agroecologia
141. Fiocruz (2019) Afinal, o Brasil é o maior consumidor de agrotóxico do mundo? | CEE Fiocruz.
http://cee.fiocruz.br/?q=node/1002
142. Feijo Fernandes CL, Martins Volcão L, Florêncio Ramires P, De Moura RR, Júnior FMRDS
(2020) Distribution of pesticides in agricultural and urban soils of Brazil: a critical review.
Environ Sci: Processes Impacts 22:256–270. https://doi.org/10.1039/C9EM00433E
143. Castelo-Grande T, Augusto PA, Monteiro P, Estevez AM, Barbosa D (2010) Remediation of
soils contaminated with pesticides: a review. Int J Environ Anal Chem 90:438–467. https://doi.
org/10.1080/03067310903374152
144. Prudente IRG, Cruz CL, de Carvalho Nascimento L, Kaiser CC, Guimarães AG (2018)
Evidence of risks of renal function reduction due to occupational exposure to agrochemicals:
a systematic review. Environ Toxicol Pharmacol 63:21–28. https://doi.org/10.1016/j.etap.
2018.08.006
145. Canal Rural (2019) FAO mostra que uso de defensivos no Brasil é menor do que em países
europeus. Canal Rural. https://www.canalrural.com.br/noticias/agricultura/fao-mostra-que-
uso-de-defensivos-no-brasil-e-menor-do-que-em-paises-europeus/
146. Jeschke P (2016) Propesticides and their use as agrochemicals. Pest Manag Sci 72(2):210–225.
https://doi.org/10.1002/ps.4170
147. Silva DE, do Nascimento JM, da Silva RTL, Juchem CF, Ruffatto K, da Silva GL, Johann L,
Corrêa LLC, Ferla NJ (2019) Impact of vineyard agrochemicals against Panonychus ulmi
(Acari: Tetranychidae) and its natural enemy, Neoseiulus californicus (Acari: Phytoseiidae) in
Brazil. Crop Prot 123:5–11. https://doi.org/10.1016/j.cropro.2019.05.014
148. Gupta S, Khanna R (2018) Agrochemicals as a potential cause of ground water pollution: a
review. Int J Chem Stud 6(3):985–990
149. Hendges C, Schiller ADP, Manfrin J, Macedo Jr EK, Gonçalves Jr AC, Stangarlin JR (2019)
Human intoxication by agrochemicals in the region of South Brazil between 1999 and 2014. J
Environ Sci Health B 54(4):219–225. https://doi.org/10.1080/03601234.2018.1550300
150. dos Santos CF, Otesbelgue A, Blochtein B (2018) The dilemma of agricultural pollination in
Brazil: beekeeping growth and insecticide use. PLoS One 13(7):e0200286. https://doi.org/10.
1371/journal.pone.0200286
151. Zoffoli HJO, do Amaral-Sobrinho NMB, Zonta E, Luisi MV, Marcon G, Tolón-Becerra A
(2013) Inputs of heavy metals due to agrochemical use in tobacco fields in Brazil’s Southern
region. Environ Monit Assess 185(3):2423–2437. https://doi.org/10.1007/s10661-012-2721-y
152. Brasil (1989) Lei no 7.802, de 11 de julho de 1989. http://www.planalto.gov.br/ccivil_03/leis/
l7802.htm
153. Souza Machado AA, Kloas W, Zarfl C, Hempel S, Rillig MC (2018) Microplastics as an
emerging threat to terrestrial ecosystems. Glob Chang Biol 24(4):1405–1416. https://doi.org/
10.1111/gcb.14020
154. Hurley RR, Nizzetto L (2018) Fate and occurrence of micro (nano) plastics in soils: knowledge
gaps and possible risks. Curr Opin Environ Sci Health 1:6–11. https://doi.org/10.1016/j.coesh.
2017.10.006
155. Horton AA, Walton A, Spurgeon DJ, Lahive E, Svendsen C (2017) Microplastics in fresh-
water and terrestrial environments: evaluating the current understanding to identify the
knowledge gaps and future research priorities. Sci Total Environ 586:127–141. https://doi.
org/10.1016/j.scitotenv.2017.01.190
126 P. T. S. Oliveira et al.

156. Kaza S, Yao L, Bhada-Tata P, Van Woerden F (2018) What a waste 2.0: a global snapshot of
solid waste management to 2050. World Bank Publications. http://hdl.handle.net/10986/3031
7
157. Kutralam-Muniasamy G, Pérez-Guevara F, Elizalde-Martínez I, Shruti VC (2020) Review of
current trends, advances and analytical challenges for microplastics contamination in Latin
America. Environ Pollut:115463. https://doi.org/10.1016/j.envpol.2020.115463
158. He D, Luo Y, Lu S, Liu M, Song Y, Lei L (2018) Microplastics in soils: analytical methods,
pollution characteristics and ecological risks. TrAC Trends Anal Chem 109:163–172. https://
doi.org/10.1016/j.trac.2018.10.006
159. Bittencourt S, Aisse MM, Serrat BM (2017) Gestão do uso agrícola do lodo de esgoto: estudo
de caso do estado do Paraná, Brasil. Engenharia Sanitaria e Ambiental 22(6):1129–1139.
https://doi.org/10.1590/S1413-41522017156260
160. ANA (2017) Agência Nacional de Águas (ANA) - Atlas Esgotos - Despoluição de bacias
hidrográficas
161. Sampaio RA, Araújo WF (2001) Importância da cobertura plástica do solo sobre o cultivo de
hortaliças. Agropecuária Técnica 22(1/2):1–12
162. Embrapa (2015) Uso do plástico na agricultura protegida. Embrapa Hortaliças. https://
Embrapa.br/busca-de-noticias/-/noticia/3230175/uso-do-plastico-na-agricultura-protegida
163. Rocha DB, Souza de Carvalho J, de Oliveira SA, Santos Rosa D (2018) A new approach for
flexible PBAT/PLA/CaCO3 films into agriculture. J Appl Polym Sci 135(35):46660. https://
doi.org/10.1002/app.46660
164. Steinmetz Z, Wollmann C, Schaefer M, Buchmann C, David J, Tröger J et al (2016) Plastic
mulching in agriculture. Trading short-term agronomic benefits for long-term soil degrada-
tion? Sci Total Environ 550:690–705. https://doi.org/10.1016/j.scitotenv.2016.01.153
165. Ivar JAS, Spengler A, Costa MF (2009) Here, there and everywhere. Small plastic fragments
and pellets on beaches of Fernando de Noronha (Equatorial Western Atlantic). Mar Pollut Bull
58(8):1236–1238. https://doi.org/10.1016/j.marpolbul.2009.05.004
166. Deuschle D, Minella JPG, Hörbe TAN, Londero AL, Schneider FJA (2019) Erosion and
hydrological response in no-tillage subjected to crop rotation intensification in southern Brazil.
Geoderma 340:157–163. https://doi.org/10.1016/j.geoderma.2019.01.010
167. OEC (2021) The observatory of economic complexity. https://oec.world/. Accessed Mar 2021
168. Brovini EM, de Deus BCT, Vilas-Boas JA, Quadra GR, Carvalho L, Mendonça RF et al
(2021) Three-bestseller pesticides in Brazil: freshwater concentrations and potential environ-
mental risks. Sci Total Environ 771:144754. https://doi.org/10.1016/j.scitotenv.2020.144754
169. Paumgartten FJ (2020) Pesticides and public health in Brazil. Curr Opin Toxicol 22:7–11.
https://doi.org/10.1016/j.cotox.2020.01.003
170. Li C, Zhou K, Qin W, Tian C, Qi M, Yan X, Han W (2019) A review on heavy metals
contamination in soil: effects, sources, and remediation techniques. Soil Sediment Contam Int
J 28(4):380–394. https://doi.org/10.1080/15320383.2019.1592108
171. Kouamé IK, Kouassi LK, Dibi B, Adou KM, Rascanu ID, Romanescu G et al (2013) Potential
groundwater pollution risks by heavy metals from agricultural soil in Songon area (Abidjan,
Côte d’Ivoire). J Environ Prot. https://doi.org/10.4236/jep.2013.412165
172. Kelepertzis E (2014) Accumulation of heavy metals in agricultural soils of Mediterranean:
insights from Argolida basin, Peloponnese, Greece. Geoderma 221:82–90. https://doi.org/10.
1016/j.geoderma.2014.01.007
173. Su C (2014) A review on heavy metal contamination in the soil worldwide: situation, impact
and remediation techniques. Environ Skept Crit 3(2):24
174. dos Santos SN, Alleoni LRF (2013) Reference values for heavy metals in soils of the Brazilian
agricultural frontier in Southwestern Amazônia. Environ Monit Assess 185:5737–5748.
https://doi.org/10.1007/s10661-012-2980-7
175. Campos ML, Pierangeli MAP, Guilherme LRG, Marques JJ, Curi N (2003) Baseline concen-
tration of heavy metals in Brazilian Latosols. Commun Soil Sci Plant Anal 34(3–4):547–557.
https://doi.org/10.1081/CSS-120017838
Agricultural Land Degradation in Brazil 127

176. Fadigas FDS, Sobrinho N, do Amaral MB, Mazur N, dos Anjos LH, Freixo AA (2006)
Proposition of reference values for natural concentration of heavy metals in Brazilian soils.
Revista Brasileira de Engenharia Agrícola e Ambiental 10(3):699–705. https://doi.org/10.
1590/S1415-43662006000300024
177. Freitas EV, Nascimento CW, Souza A, Silva FB (2013) Citric acid-assisted phytoextraction
of lead: a field experiment. Chemosphere 92(2):213–217. https://doi.org/10.1016/j.
chemosphere.2013.01.103
178. Quadros AF (2010) Os isópodos terrestres são boas ferramentas para monitorar e restaurar
áreas impactadas por metais pesados no Brasil. Oecologia Australis 14(2):569–583. https://
doi.org/10.4257/oeco.2010.1402.13
179. Andrade MGD, Melo VDF, Gabardo J, Souza LCDP, Reissmann CB (2009) Metais pesados
em solos de área de mineração e metalurgia de chumbo: I-Fitoextração. Rev Bras Ciênc Solo
33:1879–1888. https://doi.org/10.1590/S0100-06832009000600037
180. Zampieri BDB, Pinto AB, Schultz L, de Oliveira MA, de Oliveira AJFC (2016) Diversity and
distribution of heavy metal-resistant bacteria in polluted sediments of the Araça Bay, São
Sebastião (SP), and the relationship between heavy metals and organic matter concentrations.
Microb Ecol 72(3):582–594. https://doi.org/10.1007/s00248-016-0821-x
181. Webb NP, Marshall NA, Stringer LC, Reed MS, Chappell A, Herrick JE (2017) Land
degradation and climate change: building climate resilience in agriculture. Front Ecol Environ
15(8):450–459. https://doi.org/10.1002/fee.1530
182. Giannini TC, Costa WF, Cordeiro GD, Imperatriz-Fonseca VL, Saraiva AM, Biesmeijer J,
Garibaldi LA (2017) Projected climate change threatens pollinators and crop production in
Brazil. PLoS One 12(8):e0182274. https://doi.org/10.1371/journal.pone.0182274
183. Resende NC, Miranda JH, Cooke R, Chu ML, Chou SC (2019) Impacts of regional climate
change on the runoff and root water uptake in corn crops in Parana, Brazil. Agric Water Manag
221:556–565. https://doi.org/10.1016/j.agwat.2019.05.018
184. Siqueira PP, Oliveira PTS, Bressiani D, Neto AAM, Rodrigues DB (2021) Effects of climate
and land cover changes on water availability in a Brazilian Cerrado basin. J Hydrol Reg Stud
37:100931. https://doi.org/10.1016/j.ejrh.2021.100931
185. Rodrigues JA, Viola MR, Alvarenga LA, de Mello CR, Chou SC, de Oliveira VA et al (2020)
Climate change impacts under representative concentration pathway scenarios on streamflow
and droughts of basins in the Brazilian Cerrado biome. Int J Climatol 40(5):2511–2526.
https://doi.org/10.1002/joc.6347
Agriculture Land Degradation in Chile

Marcos Francos

Contents
1 Origins and Evolution of Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
2 Pest Management Impact on Soil Contamination in Agricultural Areas . . . . . . . . . . . . . . . . . . . 136
3 Soil Water Stress and Climate Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
4 Agrochemical Use and Soil Contamination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.1 Agricultural Fertilisers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.2 Microplastics Contamination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
5 Land-Use Change and Geomorphological Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6 Final Remarks and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

Abstract Soil is one of the most important and non-renewable elements of the
environment. In order to improve income and increase productivity from the soil,
various techniques (e.g. tillage) and substances (e.g. pesticides, fertilisers) have been
employed. The use of these techniques and substances increased soil degradation.
This is coupled with other socioeconomic processes that occurred in recent years,
such as the land-use change from forest areas to agricultural lands. For this reason, in
recent decades, awareness has been raised about these issues, and various regulations
have been dictated on the need to protect the environment and the soil in agricultural
areas, identify sustainable practices to implement in agricultural areas, control the
number of fertilisers and pesticides used, and protect the mosaic of territories to
avoid further degradation.

M. Francos (✉)
Department of Geography, Faculty of Geography and History, University of Salamanca,
Salamanca, Spain
Centro de Investigación de Estudios Avanzados del Maule (CIEAM), Universidad Católica del
Maule, Talca, Chile
e-mail: [email protected]

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 129
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 129–152, DOI 10.1007/698_2022_921,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 14 December 2022
130 M. Francos

Keywords Ecosystem services, Land-use change, Soil degradation, Sustainable


agriculture

1 Origins and Evolution of Agriculture

Agriculture has accompanied the existence of humans for approximately the last
12,000–10,000 years [1], with the first evidence of cultivated land being in South-
west Asia [2]. At that time, annual and perennial crops were cultivated and relied on
organic amendments after soil preparation [3]. Agriculture was an alternative to
hunting and gathering. Humans began by sowing wild plants and domesticating
animals [4, 5].
One of the earliest recorded areas of agricultural development is the Near East
[6]. Here, the first domesticated plant was found—einkorn, emmer, barley, and pea
[7]. Some authors found that lentil cultivation began around 7,500–6,500 years BC
in Southwest Asia and the Mediterranean region [8], specifically in areas that
correspond today to the territories of Iraq, Israel, Turkey, and Syria, as detailed by
Flannery [6]. The history of Castanea sativa (sweet chestnut) in some of these
countries, such as Syria, began around 8,600 BP [9]. Rice in East Asia was planted
around 8,500 BP [10], which is considered the most important crop [11, 12]. Rice
origin is not as well documented as other major cereals like wheat and barley
[6]. Evidence of early rice cultivation was found along the Indus River valley
(3,000 BC) and in Taiwan (2,500 BC) [13]. In Mesoamerica, the first crops culti-
vated were agave, beans, squash, pumpkins, amaranths, chillies, tomatoes, and
avocados, all dating from 5,000 BC [14].
In the case of South America, squash (Cucurbita pepo and C. moschata) has been
domesticated in southern Mexico and northern Peru since 10,000–9,000 BP
[12]. Critical changes in South American societies led to the transition from a
hunter-gatherer to an agricultural way of life in mountain areas in the Andes
[15]. Around 10,000 BC, agriculture and domestic animal herding began to take
place from the north to the south of Chile [16–18]. Based on the archaeological
evidence, although today maize and pumpkin are among the dominant crops in
northern Chile, around 10,000 BC, the people ate mainly wild fruits, especially carob
(Prosopis sp.) [19].
Due to the geological (volcanism) and geomorphological (steep slopes) condi-
tions in northern Chile, soils are shallow [20]; the only cultivable areas are the valley
bottoms [21]. The shallow soils in the north of Chile make cultivation areas scarce,
where only some sloping areas are deep enough to allow for cultivation and
irrigation of the terraces and provide them with water [22]. This has been used for
many crops and extends the cultivated area from the valley bottoms up the slope,
necessary for two reasons: (1) the lack of space in valley bottoms and (2) the need to
produce food. In South-central Chile, the evidence of agriculture goes back to
11,000 BP with potato cultivation. Soon after the cultivation of potatoes, other
species were domesticated, such as squash (10,000 BP) and peanuts (8,500 BP),
Agriculture Land Degradation in Chile 131

bottle gourd (8,000–7,000 BP), and corn [12]. Only one Chilean territory that
followed a different dynamic was Easter Island (Rapa Nui). With the arrival of the
first colonisers, several crops started to be cultivated, such as dryland taro (Colocasia
esculenta), yam (Dioscorea spp.), and ti (Cordyline) [23]. In the case of wine
production, the farm owners began to work and produce together, giving rise to
the greater specialisation of the existing processes [24]. Social changes caused the
conversion of small farming estates into large latifundios, about which there is a lack
of information [25]. Chile is a country with large agricultural production, ranking as
the second-largest avocado producer in the world [26]. Agriculture is the second
most relevant primary activity after mining [27]. The agriculture sector represents
8.6% of the country’s labour force and generates a gross domestic product (GDP) of
US$258.16 billion (8% of the total Chilean GDP) [28].
With the evolution of agriculture in Chile over the years, the area cultivated by
each species changed as well (Table 1). Chile’s main agricultural products are
grapes, apples, tomatoes, wheat, maize, and potatoes. In 1969, the most widespread
crops were wheat (51.44%), grapes (6.92%), oats (5.60%), potatoes (5.28%), and
maize (4.05%). Ten years later, the dominant crops were wheat (36.29%), maize
(8.45%), grapes (7.45%), dry beans (7.12%), potatoes (5.24%), and oats (5.10%). In
1989, these were wheat (37.36%), maize (8.62%), grapes (8.19%), oats (4.75%), dry
beans (4.38%), potatoes (4.33%), and rapeseed (4.23%). By 1999, the main crop was
wheat (29.05%), with grapes (12.58%), oats (6.81%), maize (6.29%), potatoes
(5.19%), and sugar beet (4.29%). The crops occupying more than 4% of the total
agricultural surface area in 2009 were wheat (23.14%), grapes (16.41%), maize
(10.57%), and oats (8.34%). Finally, in 2019, the latest official data available show
the following distribution of major crops: wheat (18.61%), grapes (16.32%), maize
(6.72%), oats (6.23%), and rapeseed (4.02%). In Chile, the cultivated area increased
from 1,444,377 ha to 1,544,223 ha between 1969 and 1979 and then decreased
to 1,446,355 ha in 1989. The agricultural area underwent an important decrease, to
1,165,405 ha, in 1999, but in 2009, it increased to 1,212,662 ha, decreasing again to
1,196,777 ha in 2019 [27].
Agriculture in Chile has been affected by a variety of environmental problems.
Some of these problems are related to crop pests [29, 30]. One of the most-used
management procedures is tree thinning, favouring crops by eliminating the com-
petition for light and water that other plant species can generate [31]. This can be
done manually to avoid direct soil compaction and degradation due to the use of
machinery [32]. In other cases, management can be used to improve crop yields by
keeping crop residues on the soil and carrying out conservation tillage [33].
This chapter focuses on the origins and evolution of agriculture in Chile to
understand its current state and the problems that occur in soil in agricultural
areas. The work is focused on examining pest and management actions, the impact
of soil water stress and climate change, the sources of soil contamination, and the
implications of land-use change in Chile’s agricultural areas and providing informa-
tion about sustainable alternatives to avoid soil and environmental degradation.
Table 1 Evolution of the cultivated area of each species in the last few decades in Chile
132

% Crop/ % Crop/ % Crop/ % Crop/ % Crop/ % Crop/


Crop (ha) in each year year year year year year
decade 1969 1969 1979 1979 1989 1989 1999 1999 2009 2009 2019 2019
Almonds 0 0.00 0 0.00 3,700 0.26 5,900 0.51 6,900 0.57 8,867 0.74
Apples 13,800 0.96 14,735 0.95 23,000 1.59 37,400 3.21 35,100 2.89 32,371 2.70
Apricots 1,100 0.08 1,600 0.10 1980 0.14 2,406 0.21 1770 0.15 664 0.06
Artichokes 2,700 0.19 2,300 0.15 1936 0.13 3,107 0.27 5,875 0.48 1,433 0.12
Asparagus 0 0.00 0 0.00 3,600 0.25 4,183 0.36 2,936 0.24 2,102 0.18
Avocados 3,500 0.24 5,400 0.35 7,945 0.55 20,181 1.73 33,500 2.76 29,224 2.44
Barley 44,350 3.07 59,770 3.87 24,590 1.70 26,502 2.27 18,513 1.53 28,605 2.39
Bast fibres 0 0.00 5,000 0.32 10,000 0.69 12,520 1.07 11,367 0.94 10,195 0.85
Beans, dry 44,019 3.05 109,990 7.12 63,400 4.38 29,058 2.49 16,709 1.38 10,248 0.86
Beans, green 11,000 0.76 13,507 0.87 9,615 0.66 5,475 0.47 2,890 0.24 2,632 0.22
Cabbages and 1,100 0.08 1,500 0.10 2,105 0.15 2,196 0.19 1733 0.14 2,261 0.19
other brassicas
Carrots and 1,520 0.11 3,200 0.21 5,275 0.36 3,538 0.30 4,638 0.38 3,952 0.33
turnips
Cauliflowers and 630 0.04 560 0.04 813 0.06 1,621 0.14 1,505 0.12 1869 0.16
broccoli
Cherries 1,200 0.08 1,800 0.12 2,965 0.20 5,288 0.45 12,500 1.03 38,392 3.21
Cherries 0 0.00 0 0.00 0 0.00 236 0.02 85 0.01 4 0.00
Chestnut 0 0.00 0 0.00 0 0.00 0 0.00 366 0.03 1,263 0.11
Chick peas 8,700 0.60 16,810 1.09 7,830 0.54 2,266 0.19 1887 0.16 897 0.07
Chillies and 1980 0.14 2,100 0.14 2,734 0.19 3,871 0.33 1,489 0.12 763 0.06
peppers
Cranberries 0 0.00 0 0.00 0 0.00 30 0.00 6,775 0.56 18,373 1.54
Cucumbers and 880 0.06 950 0.06 1,290 0.09 1,000 0.09 1,407 0.12 1,682 0.14
gherkins
M. Francos
Currants 0 0.00 0 0.00 0 0.00 3 0.00 4 0.00 1 0.00
Figs 0 0.00 0 0.00 0 0.00 25 0.00 99 0.01 65 0.01
Flax fibre and tow 2000 0.14 2000 0.13 2000 0.14 2076 0.18 2,597 0.21 2,837 0.24
Fruit, freshness 695 0.05 1,100 0.07 2000 0.14 2,273 0.20 2,566 0.21 2,854 0.24
Garlic 1800 0.12 2,700 0.17 2,301 0.16 3,142 0.27 1,253 0.10 1,556 0.13
Grapefruit (inc. 0 0.00 0 0.00 0 0.00 352 0.03 291 0.02 272 0.02
pomelos)
Grapes 100,000 6.92 115,000 7.45 118,486 8.19 146,562 12.58 199,000 16.41 195,357 16.32
Hazelnuts, with 0 0.00 0 0.00 0 0.00 1711 0.15 6,678 0.55 24,437 2.04
shell
Hemp tow waste 3,600 0.25 3,700 0.24 4,200 0.29 4,200 0.36 4,600 0.38 4,381 0.37
Agriculture Land Degradation in Chile

Hempseed 3,600 0.25 1,000 0.06 1,000 0.07 1,083 0.09 1,602 0.13 3,323 0.28
Kiwi fruit 0 0.00 0 0.00 11,810 0.82 7,865 0.67 10,800 0.89 7,595 0.63
Lemons and 5,400 0.37 6,375 0.41 5,800 0.40 7,414 0.64 7,700 0.63 6,655 0.56
limes
Lentils 12,645 0.88 50,360 3.26 14,690 1.02 3,170 0.27 955 0.08 1731 0.14
Lettuce and 1,600 0.11 3,500 0.23 4,184 0.29 5,991 0.51 7,357 0.61 6,476 0.54
chicory
Linseed 810 0.06 1,000 0.06 1,000 0.07 1,071 0.09 1,519 0.13 4,157 0.35
Lupins 0 0.00 2000 0.13 8,570 0.59 18,724 1.61 10,283 0.85 21,280 1.78
Maize 58,440 4.05 130,410 8.45 124,650 8.62 73,284 6.29 128,211 10.57 80,428 6.72
Maize, green 9,700 0.67 9,900 0.64 11,402 0.79 14,774 1.27 11,458 0.94 10,151 0.85
Melons, other 5,300 0.37 6,600 0.43 3,705 0.26 3,733 0.32 3,130 0.26 3,091 0.26
(inc. cantaloupes)
Oats 80,830 5.60 78,720 5.10 68,690 4.75 79,402 6.81 101,101 8.34 74,617 6.23
Olives 0 0.00 0 0.00 3,400 0.24 4,481 0.38 12,000 0.99 21,951 1.83
Onions 3,500 0.24 8,800 0.57 10,320 0.71 5,892 0.51 6,144 0.51 7,919 0.66
Oranges 4,000 0.28 4,983 0.32 6,270 0.43 7,237 0.62 7,500 0.62 6,244 0.52
133

(continued)
Table 1 (continued)
134

% Crop/ % Crop/ % Crop/ % Crop/ % Crop/ % Crop/


Crop (ha) in each year year year year year year
decade 1969 1969 1979 1979 1989 1989 1999 1999 2009 2009 2019 2019
Papayas 112 0.01 60 0.00 310 0.02 375 0.03 183 0.02 151 0.01
Peaches and 14,000 0.97 13,800 0.89 16,700 1.15 18,003 1.54 21,000 1.73 15,651 1.31
nectarines
Pears 2,900 0.20 2,750 0.18 14,520 1.00 10,675 0.92 6,600 0.54 7,387 0.62
Peas, dry 10,880 0.75 16,660 1.08 6,550 0.45 1813 0.16 2,752 0.23 1,173 0.10
Peas, green 7,500 0.52 4,100 0.27 6,183 0.43 4,905 0.42 2,969 0.24 1743 0.15
Persimmons 0 0.00 0 0.00 0 0.00 96 0.01 129 0.01 112 0.01
Plums and sloes 2,600 0.18 2,858 0.19 8,435 0.58 13,055 1.12 18,500 1.53 17,811 1.49
Potatoes 76,215 5.28 80,930 5.24 62,680 4.33 60,465 5.19 45,078 3.72 41,811 3.49
Pumpkins, 5,220 0.36 6,850 0.44 5,448 0.38 5,038 0.43 5,498 0.45 3,487 0.29
squash, and
gourds
Quinces 0 0.00 0 0.00 0 0.00 491 0.04 379 0.03 318 0.03
Rapeseed 48,380 3.35 53,900 3.49 61,120 4.23 31,995 2.75 25,135 2.07 48,166 4.02
Raspberries 0 0.00 0 0.00 0 0.00 4,630 0.40 4,559 0.38 3,888 0.32
Rice 16,190 1.12 47,070 3.05 42,990 2.97 14,696 1.26 23,680 1.95 26,242 2.19
Rye 8,396 0.58 7,370 0.48 2,870 0.20 1,360 0.12 2,135 0.18 568 0.05
Soybeans 1,000 0.07 1800 0.12 0 0.00 0 0.00 0 0.00 0 0.00
Strawberries 0 0.00 0 0.00 820 0.06 768 0.07 1,518 0.13 1,064 0.09
Sugar beet 27,290 1.89 14,874 0.96 51,600 3.57 50,053 4.29 12,869 1.06 12,919 1.08
Sunflower seed 24,630 1.71 21,660 1.40 15,000 1.04 2,929 0.25 4,355 0.36 3,228 0.27
Sweet potatoes 1,000 0.07 1,000 0.06 1,000 0.07 1,077 0.09 1,303 0.11 1,410 0.12
Tangerines, man- 0 0.00 0 0.00 0 0.00 236 0.02 2,990 0.25 7,797 0.65
darins,
clementines,
M. Francos

satsumas
Tobacco, 1920 0.13 2,981 0.19 3,460 0.24 3,959 0.34 1,652 0.14 2,396 0.20
unmanufactured
Tomatoes 9,500 0.66 16,000 1.04 12,185 0.84 18,968 1.63 11,562 0.95 15,202 1.27
Triticale 0 0.00 0 0.00 0 0.00 0 0.00 17,907 1.48 27,023 2.26
Vegetables, 4,900 0.34 6,020 0.39 6,100 0.42 6,161 0.53 6,791 0.56 5,762 0.48
freshness
Vegetables, 2000 0.14 3,500 0.23 1,495 0.10 2,339 0.20 1922 0.16 1869 0.16
leguminousnes
Walnuts, with 0 0.00 0 0.00 5,600 0.39 7,565 0.65 12,600 1.04 40,801 3.41
shell
Watermelons 6,300 0.44 12,200 0.79 3,743 0.26 3,927 0.34 3,159 0.26 2,918 0.24
Agriculture Land Degradation in Chile

Wheat 743,045 51.44 560,470 36.29 540,290 37.36 338,583 29.05 280,644 23.14 222,705 18.61
Total in Chile 1,444,377 100 1,544,223 100 1,446,355 100 1,165,405 100 1,212,662 100 1,196,777 100
Source: Data extracted from FAOSTAT in the FAO (Food and Agriculture Organization) website (http://www.fao.org/home/en/). Consulted on 01/04/2021
135
136 M. Francos

2 Pest Management Impact on Soil Contamination


in Agricultural Areas

Pests are one of the greatest risks to which agricultural areas are exposed, and
landowners can lose their crops. Chemical pesticides are used to prevent pests,
remaining in the soil for long periods. Pesticides are a major source of contamination
worldwide [34]. Several studies have analysed the influence of pests in agricultural
areas of Chile [30, 35–37], but none have focused on how pest control treatments
contaminated the soil but instead evaluated the effect of pests on crops and the
effectiveness of such treatments (Fig. 1).

Fig. 1 Density of studies concerning agriculture and soil in Chile between 2010 and 2021
published in English. (a) Studies concerning agricultural pest and management strategies in Chile
between 2010 and 2021. Data source: scientific papers published mainly in English found by the
Google Scholar academic search engine for the 2010–2021 period using the following criteria in the
title: “pest” OR “management” AND “soil” OR “agriculture” AND “Chile” [38–40]; (b) Studies
concerning water stress and climate change in agriculture in Chile between 2010 and 2021. Data
source: scientific papers published mainly in English found by the Google Scholar academic search
engine for the 2010–2021 period using the following criteria in the title: “water” OR “climate”
AND “soil” OR “agriculture” AND “Chile” [33, 41–62]; (c) Studies concerning agrochemical uses
and soil contamination in agriculture in Chile between 2010 and 2021. Data source: scientific papers
published mainly in English found by the Google Scholar academic search engine for the
2010–2021 period using the following criteria in the title: “fertiliser” OR “plastic” OR “contami-
nation” AND “soil” OR “agriculture” AND “Chile” [63, 64]; (d) Studies concerning land-use
change and geomorphological processes in agriculture in Chile between 2010 and 2021. Data
source: scientific papers published mainly in English found by the Google Scholar academic search
engine for the 2010–2021 period using the following criteria in the title: “land-use” OR “geomor-
phology” AND “soil” OR “agriculture” AND “Chile” [65–71]
Agriculture Land Degradation in Chile 137

Subsoiling has proved to be a good management strategy to protect crops against


degradation. In some areas of Chile, most precipitation is concentrated in just a few
days during the wet months of the year. In these cases, subsoiling averts soil erosion
due to the accumulation of water in the subsoil, as it exceeds its saturation point. The
effectiveness of subsoiling is lost over time, being effective for approximately
4 years, making it necessary to reapply the technique [72]. In other southern
agricultural areas of Chile, conservation tillage has been used to mitigate and
rehabilitate dryland areas and avoid land degradation. Brunel-Saldias et al. [73]
compared different treatments and their impacts on soil physical properties, finding
that the best management was a zero tillage system with subsoiling. This manage-
ment method consists of using a no-till seeder and a subsoiler chisel plough that can
reach a 40-cm depth. The benefits of this practice included an increment in pore size,
crop water availability under water stress conditions, and a better root system in the
topmost centimetres with greater root exploration at deeper depths. Climent et al.
[74] observed pesticide residues in surface waters of the Cachapoal River basin,
central Chile. While this study did not focus on the impact of pesticides in soils, the
detection of pesticides in this waterway, and the use of river water for irrigation, is
causing fertilisers to reach the soils in crop fields and consequently soil
contamination.

3 Soil Water Stress and Climate Change

Water stress and variable water availability in agricultural areas can lead to land
degradation and the death of crops. In most cases, water stress leads to differences in
the natural vegetation [75] and, consequently, productivity. Different factors can
produce water stress, the most important being climate change [76], which decreases
available water resources [77]. Some studies have looked for ways to reduce water
stress [78] by quantifying the damage it causes and supporting the need for public
policies that not only incentivise productivity or provide subsidies to alleviate the
damage of water stress but also work to solve the problem from the bottom up so that
the problem in question is much less likely to occur.
Other authors have suggested using agricultural conservation systems to improve
agricultural productivity under water stress conditions [79]. In this system, there is
zero tillage, and the crop residues are deposited under soil cover. The relationship
between climate change and the reduction in water availability for crops is a topic of
study that has been widely addressed worldwide [75, 80–82]. This has become
particularly relevant in recent years as the inevitable impacts of climate change on
agricultural areas become visible [76, 83]. For this reason, many authors have not
only been raising the question of water scarcity or establishing models to predict the
situation in a few years but have also been looking for management alternatives or
crops that are much more adapted to the new conditions of the future [84, 85]. Despite
these attempts, climate change is universally agreed that it irreversibly affects water
availability, leading to water stress [86].
138 M. Francos

The close link between climate change and decreased water availability in
agricultural areas means that alternatives, such as groundwater pumping, must be
sought to avoid land degradation [87]. In the Maule region, several studies have been
carried out in which, based on evidence of a decrease in water availability due to
climate change, various alternatives have been sought to avoid land degradation. The
most important of these studies is the one carried out by Letelier et al. [41], in which
they raised the question of using water from the Atlantic Ocean to irrigate crop fields.
Other studies, carried out in the same region in central Chile, have pointed to the
high vulnerability of these agricultural areas to climate change and water stress that
lead to land degradation in agricultural areas [42]. This will inevitably lead to
changes in land use, causing policy-makers to intervene, firstly to avoid the deteri-
oration and death of agricultural fields by prioritising existing resources and sec-
ondly to avert environmental degradation by ensuring a satisfactory transition during
land-use change [88] (Fig. 1).

4 Agrochemical Use and Soil Contamination

4.1 Agricultural Fertilisers

Pesticides are used worldwide to protect crops [89]. The excessive and uncontrolled
use of these pesticides leads to the contamination of ecosystems and abiotic factors
(e.g. watersheds, soils) [34]. The proliferation in the use of fertilisers has occurred in
recent decades due to the increase in the world’s population to meet the growing
demand [90]. Fertiliser use in agriculture is a topic of study that has been widely
analysed worldwide, as have remediation treatments [91]. According to Zhou and
Wang [92], agrochemicals increase cadmium concentrations and other metals
(nickel, cobalt, copper, barium, and lead) in agricultural soils [93, 94]. The ingestion
of food plants cultivated in contaminated areas can endanger human health through
bioaccumulation (e.g. radish, garlic, barley and wheat) [95]. The pesticides and
fertilisers used in agriculture remained in the environment for a long time, contam-
inating water resources. Previous works analysed how some agrochemicals that are
not listed as pesticides affect the soil and have an important impact on the biota, and
thus on the ecosystem as a whole, causing insecticidal side effects and the death of
arthropods [96]. In addition, heavy metal contamination affects large areas under
cultivation in rural areas and has a major impact on urban gardens, affecting the
health of the consumers of these products [97].
Chile is not exempt from this contamination (Fig. 1), being one of four countries
(along with China, Australia, and Guatemala) where pollution from pesticides
intersects the habitat of at least one endangered or critically endangered amphibian
species [34]. For this reason, the use of agrochemicals in agricultural areas is a topic
of study that involves professionals from different research areas because it gener-
ates an important impact on the different components of habitat and the living beings
that inhabit it. In many cases, these studies have focused on the visible part of the
Agriculture Land Degradation in Chile 139

ecosystems, forgetting one of the most important elements – the one that takes the
longest time to recover after a disturbance such as contamination, the soil [74]. The
problem of agricultural soil contamination in Chile is a topical issue that must be
given the importance it deserves, although not many papers have been published on
this issue so far. It is mainly caused by high concentrations of copper, cadmium,
lead, and zinc and metalloids, such as arsenic. Alekseev and Abakumov [98] pointed
out that the retention of these metals and compounds in soils occurs through
mechanisms such as complexes forming with organic matter or occlusion within
organic matter. Therefore, soil contamination studies should focus on these mech-
anisms. Alekseev and Abakumov [98] analysed heavy metal contamination in
central Chile, finding that soils with higher amounts of the organic matter showed
higher abundances of aromatics and lower abundances of aliphatic matter, conclud-
ing that these soils had a higher nitrogen content and that they retained heavy metals
and contaminating compounds more effectively, which made the soils more
contaminated.
Vega et al. [99] conducted a comprehensive review of perchlorate contamination
in Chile and suggested possible solutions. One of the main sources of soil contam-
ination in Chile is saltpetre, found in the Atacama Desert. This saltpetre is commonly
used in Chile as a nitrogen fertiliser and can be found in high concentrations in water,
fruit, and wine, all of which can affect human health. Several studies have shown the
harmful impact of perchlorate on soils in agricultural areas [100]. Calderón et al.
[100] observed the impact of perchlorate contamination caused by nitrogen
fertilisers on agricultural soils cultivated with Lactuca sativa (lettuce). In these
areas, as the authors pointed out, one of the agronomic practices that produce the
greatest accumulation of fertilisers in the soil is fertigation, which is necessary
because it constitutes an important source of perchlorate in soils. Reyes et al.
[101] and Reyes et al. [102] analysed the impact of improper waste disposal
practices on the heavy metal contamination of agricultural and urban soils in
northern Chile. The authors concluded a need to develop intervention limits and
guidance for environmental legislation to restrict these pollutants in Chilean soils. In
recent years, some authors have highlighted the influence that mining has on the
contamination of agricultural areas in Chile [103, 104]. Chile is one of the countries
where mining accounts for a relatively high percentage of the GDP and is one of the
main economic resources in the central and northern parts of the country. The
existence of mining areas adjacent to agricultural fields means that, in many cases,
the soils of agricultural fields are affected by copper contamination [105]. Stowhas
et al. [105] suggested that the application of Zn could be a good alternative to deal
with this copper contamination through fixation by microorganisms. Chile is the
world’s leading producer of copper, rhenium, natural nitrates, lithium, and iodine
[106]. It also stands out for its participation in the production of molybdenum, silver,
and gold. For this reason, mining occupies large land areas, and agricultural areas
must develop methods or techniques to combat the effects of this economic activity.
In Chile, pesticides are often implicated in reported cases of acute poisoning,
mainly due to the direct exposure of workers in the agricultural sector to these
pollutants [106]. Ramírez-Santana et al. [107] highlighted the need to implement
140 M. Francos

public policies that address environmental pollution in a preventive manner. They


pointed out the impact of organophosphate and carbamate pesticides on cholines-
terase inhibition and cognitive impairment. This impact of pollution on the human
organism has been studied by several authors, including Pancetti et al. [108], in
northern Chile, where they concluded that pesticides seriously damaged the health of
agricultural workers and the surrounding population by affecting their cognitive
areas of attention, praxis, language/vocabulary, and executive function.

4.2 Microplastics Contamination

Microplastics pollution is a global issue due to its ecological and socioeconomic


impacts on the environment [109]. Microplastics and soil contamination is a topic of
study that has attracted broad interest from many parts of the world [110], and over
the years, it has been recognised as an increasing threat in agricultural areas
[111]. Zhang et al. [112] summarised the input and output of microplastics in soil.
Plastics reach the soil surface due to poor water treatment and excessive sludge
discharge, producing soil contamination [113]. Other authors have focused on the
accumulation of microplastics in the soil, mainly from atmospheric deposition [114],
the agricultural application of organic fertilisers, or plastic mulching [115]. The
presence of microplastics is considered to be a vector of organic pollutants in the
environment. Klingelhöfer et al. [116] considered the solution to this would be to
include the plastics market in a circular economic system that involved scientists,
policy-makers, and funders to avoid the pollution of soils and agricultural areas and
consequently the degradation of the environment. The use of plastics has increased
since World War II, and is expected to peak in the next 20 years, according to authors
such as Paziencia and De Lucia [117], by population growth and the need to increase
productivity in these areas. Paziencia and De Lucia [117] have suggested that the
solution to microplastic pollution in soils is multiscale. It should take into account
the size of the farm property, with the most appropriate plastic management mea-
sures being taken, while tax credits and paybacks should be applied to large farms to
try to reduce their use of plastics. Microplastic accumulation in the soil promotes
changes in the soil structure, modifying the soil pH and pore structure [118]. The
degradation of plastics that reach the soil surface varies from 20 to more than
1,000 years. This makes this a long-term problem that must be tackled immediately
with public policies and social awareness in order to decrease, and ultimately
prevent, this type of pollution in the short to long term [119]. The degradation of
plastics depends on physical, chemical, and biological processes, so the accumula-
tion and decomposition of microplastics in the soil vary in each ecosystem
[120, 121]. There is only one paper on this involving Chile, in which different
soils were compared, and the impacts of microplastics contamination in soil aggre-
gates and soil matrices were assessed, but this was not based on agricultural areas
[63]. There are no published studies on soil contamination by microplastics in
agricultural areas in Chile, but because the vast majority of owners, technicians,
Agriculture Land Degradation in Chile 141

and scientists have dedicated themselves to evaluating and improving agricultural


areas’ productivity, without considering the damage that this type of contamination
produces in these areas (Fig. 1). At the international level, Napper and Thompsom
[122] have pointed to the need for further studies on the degradation of microplastics
and their environmental consequences.

5 Land-Use Change and Geomorphological Processes

There has been some interest in the impact of land-use change on soil degradation.
The impact of land-use change is mainly determined by differences in general soil
management practices and the cropping system [123]. Historical land-use changes
have led to changed soil characteristics [124]. One of the most internationally proven
effects is soil loss and degradation due to erosion. Ferreira et al. [125] assessed the
impact of irrigation on the risk of erosion in a cultivated area and found that the areas
where erosion was most severe were the cultivated areas occupied by olive trees and
orchards. They observed how the application of Good Agricultural and Environ-
mental Conditions conservation practices has substantially reduced soil loss while
generating increased soil organic carbon stock, making these practices very useful
for preventing soil erosion and environmental degradation in agricultural areas.
In recent years, land-use change, and its impact on agricultural areas, has become
more relevant (Fig. 1). Studies on this have mainly been carried out where areas of
special natural value are affected. The extension of, and increase in, cultivated areas,
in many cases, has caused native forests to give way to shifting cultivation, leading
to a decrease in water and sediment quality in these areas [126]. Hernández et al. [65]
focused on land-use change from native forest to agricultural land of different types.
They studied the degradation of native forest and its conversion to annual and
perennial crops, which led to a profound alteration in soil quality. Another study,
carried out in southern Chile, used satellite imagery and multi-criteria analysis to
illustrate significant changes in land use over the last three decades, with many areas
changing from forest to plantations [127]. This change in land use due to the intrinsic
properties of plantations negatively affects the soil system, with monoculture crops
affecting soil properties and soil biodiversity. The expansion of agricultural areas
(in addition to mining and urbanisation), and the concomitant changes in land use,
has been studied by Montoya-Tangarife et al. [128]. The authors analysed how this
land-use change has affected ecosystem services in central Chile. Agricultural areas
have suffered a significant decline in the last few decades due to urbanisation in the
peri-urban areas of large Chilean cities, such as Santiago and Valparaiso. With the
passage of time and demographic pressure, forested areas have been disappearing in
order to supply the new residents through local consumption. Montoya-Tangarife
et al. [128] observed that this change in land use did not significantly affect large-
scale ecosystem services at the environmental level.
The land-use change from forest to plantation has also been studied by
Nahuelhual et al. [129] in terms of the expansion of crops or timber plantations
142 M. Francos

and the processes of deforestation and reforestation that have taken place in recent
decades for economic purposes. The described land-use changes produce losses in
ecosystem services at the environmental level, pointing out the need to conserve
native forest landscapes by retaining patches of native forest among a variety of
heterogeneous land uses [130]. These authors reported a significant loss of biodi-
versity and a negative impact on the soil. A study carried out by Braun et al. [131] is
one of those few. They examined land-use change from forested areas to plantations
for economic gain and the potential erosion risk associated with these areas. The
study revealed how plantations associated show higher erosion rates and soil deg-
radation than previously native forests, with corrective measures being needed at
many points to curb the erosion. Estimates by authors such as Díaz et al. [132]
indicate that the most vulnerable areas for future land-use changes are agricultural
areas giving way to timber plantations, even in mountainous areas, so that protecting
agricultural areas and sustainable agriculture is a measure that should be taken
immediately.

6 Final Remarks and Conclusions

Crops in Chile have diversified over several decades, with many different species
having been planted. This means that those diseases that affect only one type of crop
and thus could damage the entire production system do not pose a significant threat
in Chile due to the percentage distribution spread among diverse crops, so the
production system is maintained. The increase, over the years, in the fragility,
profitability, and quality of the crops has led to greater care being taken of them
and greater protection directed towards agricultural lands. Considering the incidence
of pests in Chile as a whole and knowing the distribution of the various crops
throughout the country, it can be understood that the areas most affected by these
pests are the central and southern regions of Chile, which is where most of the
agricultural fields are concentrated.
Fertiliser depletion must be dealt with proactively, not only after the fact. After
the fact, severe corrective measures must be established, and transgressors who
cause such pollution must be subject to severe sanctions to force them to rectify
the environmental damage. In the case of fertilisers in Chile, studies need to be
carried out not only on the impact of fertilisers on soil but also on sediments, surface
water and groundwater, and foodstuffs. Studies are also required into ways of
mitigating or averting damage to human health. For example, finding new
perchlorate-reducing microorganisms in the Atacama Desert opens a door that
might allow biological treatments to deal with soil contamination and environmental
degradation. Studies on the microplastics contamination of soil in agricultural areas
in Chile are non-existent, so there is a lack of knowledge on this topic that needs to
be addressed. General plastics degradation studies should be carried out, and
microplastics should be included, in terms of their contribution to agricultural soil
Agriculture Land Degradation in Chile 143

contamination and degradation at the national and international level, so that appro-
priate policies and corrective measures can be developed.
Most studies on land-use change and its impact on soil characteristics are based
on the conversion from native forest to tree plantations or agricultural lands. In
Chile, such studies have mainly been carried out on the central-southern region
because the Atacama Desert covers the country’s northern half. The conversion of
forest to agricultural land is likely due to the difficulties associated with cultivating in
these challenging areas. All changes in land use must be controlled and regulated to
avoid soil degradation. In many cases, natural forest areas are converted to exotic
plantations or agricultural areas due to market pressures and the human desire to
obtain (more) income from the environment. It is necessary to carry out land-use
planning to generate a landscape mosaic of different land uses to protect natural
areas. Therefore, the objective is not to limit the connectivity of biodiversity.
There has not been enough space for conversation within the national context in
Chile on some fundamental issues of agricultural land degradation that need to be
addressed in greater depth. Due to the low number of articles written in Chile and the
difficulties in establishing firm conclusions on soil compaction, overgrazing, salin-
ity, and slash and burn agriculture. These have not been incorporated as subsections
in the chapter, but it is noted that there are some studies on these topics carried out in
Chile and that the study of these issues has been gaining relevance in recent years
[70, 133–137].
The areas where a greater number of studies published internationally are dis-
tributed in Chile depend on various causes such as a) the existence of a greater
agricultural extension favoured by more suitable climatic conditions for the devel-
opment of different crops and b) greater concern for this topic of study, mainly
caused by greater development of the universities in the area in this problem (Fig. 1).
Agricultural areas are an environment where soil contamination and degradation is
very widespread. In addition, these areas are an essential source of resources and
food for humans, and the health of agricultural lands is directly related to the health
of the food produced there and that people consume. Therefore, protecting the soil of
these areas from pollution, and developing sustainable treatments and production
systems for the benefit of all, should be everyone’s task.

Acknowledgements The author also wishes to thank Cambridge Proofreading & Editing LLC for
the English revision of this chapter and to Oscar Corvacho-Ganahín for his help with the carto-
graphic design.

Conflicts of interest The author declares no conflicts of interest.

References

1. Lu TLD (2005) The origin and dispersal of agriculture and human diaspora in East Asia. In: Lu
TLD (ed) The peopling of East Asia. Routledge, London, pp 75–86
144 M. Francos

2. Wallace M, Jones G, Charles M, Forster E, Stillman E, Bonhomme V, Livarda A, Osborne CP,


Rees M, Frenck G, Preece C (2019) Re-analysis of archaeobotanical remains from pre-and
early agricultural sites provides no evidence for a narrowing of the wild plant food spectrum
during the origins of agriculture in southwest Asia. Veg Hist Archaeobotany 28(4):449–463
3. Iriarte J, Elliott S, Maezumi SY, Alves D, Gonda R, Robinson M, de Souza JG, Watling J,
Handley J (2020) The origins of Amazonian landscapes: plant cultivation, domestication and
the spread of food production in tropical South America. Quaternary Sci Rev 248:106582
4. Snir A, Nadel D, Groman-Yaroslavski I, Melamed Y, Sternberg M, Bar-Yosef O, Weiss E
(2015) The origin of cultivation and proto-weeds, long before Neolithic farming. PLoS One
10(7):e0131422
5. Jones G, Kluyver T, Preece C, Swarbrick J, Forster E, Wallace M, Charles M, Rees M,
Osborne CP (2020) The origins of agriculture: intentions and consequences. J Archaeol Sci
125:105290
6. Flannery KV (1973) The origin of agriculture. Ann Rev Anthropol 2:271–310
7. Harlan JR (1992) Crops and man. American Society of Agronomy, Madison
8. Cokkizgin A, Shtaya MJ (2013) Lentil: origin, cultivation techniques, utilisation and advances
in transformation. Agr Sci 1(1):55–62
9. Conedera M, Krebs P, Tinner W, Pradella M, Torriani D (2004) The cultivation of Castanea
sativa (Mill.) in Europe, from its origin to its diffusion on a continental scale. Veget Hist
Archaeobot 13:161–179
10. Lu TLD (1999) The transition from foraging to farming and the origin of agriculture in China.
British Archaeological Reports Limited
11. Khush GS (1997) Origin, dispersal, cultivation and variation of rice. Plant Mol Biol 35:25–34
12. Reed CA (2011) Origins of agriculture. Walter de Gruyter, Berlin
13. Chang KC (1970) The beginnings of agriculture in the Far East. Antiquity 44:75–85
14. Flannery KV (1968) Archeological systems theory and early Mesoamerica. In: Meggers B
(ed) Anthropological archeology in the Americas. Anthropological Society, Washington, pp
67–87
15. Santana-Sagredo F, Uribe M, Herrera MJ, Retamal R, Flores S (2015) Dietary practices in
ancient populations from northern Chile during the transition to agriculture (Tarapacá region,
1000 BC–AD 900). Am J Phys Anthropol 158(4):751–758
16. Atencio LN (1982) Temprana emergencia de sedentarismo en el desierto chileno: proyecto
Caserones. Chungara:80–122
17. Muñoz I (2004) El periodo Formativo en los valles del norte de Chile y sur de Perú: Nuevas
evidencias y comentarios. Chungará 36:213–225
18. Watson JT, Arriaza B, Standen V, Muñoz Ovalle I (2013) Tooth wear related to marine
foraging, agro-pastoralism and the formative transition on the northern Chilean coast. Int J
Osteoarchaeol 23(3):287–302
19. Knudson KJ, Torres-Rouff C (2009) Investigating cultural heterogeneity in San Pedro de
Atacama, northern Chile, through biogeochemistry and bioarchaeology. Am J Phys Anthropol
138(4):473–485
20. Köppen W (1900) Versuch einer Klassifikation der Klimate, vorzugsweise nach ihren
Beziehungen zur Pflanzenwelt. Geogr Zeitschrift 6(593–611):657–679
21. Wright ACS (1963) The soil process and the evolution of agriculture in northern Chile. Pac
Viewp 4(1):65–74
22. Sandor JA, Huckleberry G, Hayashida FM, Parcero-Oubiña C, Salazar D, Troncoso A, Ferro-
Vázquez C (2021) Soils in ancient irrigated agricultural terraces in the Atacama Desert, Chile.
Geoarcheology:1–24. https://doi.org/10.1002/gea.21834
23. Stevenson CM, Jackson TL, Mieth A, Bork HR, Ladefoged TN (2006) Prehistoric and early
historic agriculture at Maunga Orito, Easter Island (Rapa Nui), Chile. Antiquity 80(310):919
24. Hellin J, Lundy M, Meijer M (2009) Farmer organisation, collective action and market access
in Meso-America. Food Policy 34(1):16–22
Agriculture Land Degradation in Chile 145

25. Robles C (2020) The agrarian historiography of Chile: Foundational interpretations, conven-
tional reiterations, and critical revisionism. Hist Agrar 8:93–122
26. Monzón VH, Avendaño-Soto P, Araujo RO, Garrido R, Mesquita-Neto JN (2020) Avocado
crops as a floral resource for native bees of Chile. Rev Chil Hist Nat 93:1–7
27. Oficina de Estudios y Políticas Agrarias (ODEPA) (2017) Panorama de la agricultura chilena.
Ministerio de Agricultura, Santiago. https://www.odepa.gob.cl/wp-content/uploads/2017/12/
panoramaFinal20102017Web.pdf
28. Muñoz-Schick C, Mattar Bader C, Neira Roa R, Mora González M, Espinoza J, Seguel
Seguel O, Salazar Guerrero O, Fuster Gómez R, Lizana A, Cofré C, Pinheiro A, Rodríguez
L (2017) Sustainable agriculture and healthy food in Chile. In: Asenjo J, McNeil J (eds)
Challenges and opportunities for food and nutrition security in the Americas, the view of
academies of science. Federal Ministry of Education and Research, The Inter-American
Network of Academies of Sciences, Leopoldina National Akademie der Wissenschaften and
The InterAcademy Partnership. http://repositorio.uchile.cl/handle/2250/147838, pp 190–211
29. Curkovic T, Araya J, Canales C, Medina A (2005) Evaluation of two agriculture detergents as
control alternatives for green peach aphid and twospotted spidermite, two pests affecting peach
orchards in Chile. In: VI international peach symposium, vol 713, pp 405–408
30. Rodríguez-San Pedro A, Allendes JL, Beltrán CA, Chaperon PN, Saldarriaga-Córdoba MM,
Silva AX, Grez AA (2020) Quantifying ecological and economic value of pest control services
provided by bats in a vineyard landscape of central Chile. Agric Ecosyst Environ 302:107063
31. Cruz GE, Rodriguez FA, Tapia PA, Bown HE (2018) Respuestas en crecimiento después de
un raleo con árboles futuro y un raleo por lo bajo en un bosque secundario de Nothofagus
pumilio, en Tierra del Fuego, Chile. Cienc Investig Agrar 45(3):263–276
32. Von Bennewitz E, Sanhueza S, Elorriaga A (2010) Effect of different crop load management
strategies on fruit production and quality of sweet cherries (Prunus avium L.) ‘Lapins’ in
Central Chile. J Fruit Ornam Plant Res 18(1):51–57
33. Martínez I, Ovalle C, Del Pozo A, Uribe H, Valderrama N, Prat C, Sandoval M, Fernández F,
Zagal E (2011) Influence of conservation tillage and soil water content on crop yield in dryland
compacted Alfisol of Central Chile. Chil J Agr Res 71(4):615–622
34. Tang FH, Lenzen M, McBratney A, Maggi F (2021) Risk of pesticide pollution at the global
scale. Nat Geosci 14:206–210
35. Aroca GE, Ramírez ME, Robotham H, Avila M (2020) Morphological and reproductive
studies on the green filamentous pest Rhizoclonium-like affecting Agarophyton chilensis
commercial farms in southern Chile. Aquat Bot 167:103291
36. Figueroa CC, Simon JC, Le Gallic JF, Prunier-Leterme N, Briones LM, Dedryver CA,
Niemeyer HM (2005) Genetic structure and clonal diversity of an introduced pest in Chile,
the cereal aphid Sitobion avenae. Heredity 95(1):24–33
37. Larraín S (2002) Incidencia de insectos y ácaros plagas en pepino dulce (Solanum muricatum
Ait.) cultivado en la IV región, Chile. Agr Téc 62(1):15–26
38. Ellena M, Sandoval P, Gonzalez A, Montenegro A, Jequier J, Contreras M, Azocar G (2012)
Preliminary observations on the effects of soil management techniques on hazelnut growing in
the Gorbea area, in the South of Chile. In: VIII international congress on hazelnut, vol 1052, pp
225–230
39. Ortega-Blu R, Molina-Roco M (2016) Evaluation of vegetation indices and apparent soil
electrical conductivity for site-specific vineyard management in Chile. Precis Agric 17(4):
434–450
40. Ordóñez I, López IF, Kemp PD, Descalzi CA, Horn R, Zúñiga F, Dörner J (2018) Effect of
pasture improvement managements on physical properties and water content dynamics of a
volcanic ash soil in southern Chile. Soil Tillage Res 178:55–64
41. Letelier L, Gaete-Eastman C, Peñailillo P, Moya-León MA, Herrera R (2020) Southern species
from the biodiversity hotspot of central Chile: a source of color, aroma, and metabolites for
global agriculture and food industry in a scenario of climate change. Front Plant Sci 11:1002
146 M. Francos

42. Fernández FJ, Blanco M, Ponce RD, Vásquez-Lavín F, Roco L (2019) Implications of climate
change for semi-arid dualistic agriculture: a case study in Central Chile. Reg Environ Chang
19(1):89–100
43. Fuentealba A, Duran L, Morales NS (2021) The impact of forest science in Chile: history,
contribution, and challenges. Can J For Res 51(6):753–765
44. Vicuña S, McPhee J, Garreaud RD (2012) Agriculture vulnerability to climate change in a
snowmelt-driven basin in semiarid Chile. J Water Res Plan Manag 138(5):431–441
45. Meza FJ, Wilks DS, Gurovich L, Bambach N (2012) Impacts of climate change on irrigated
agriculture in the Maipo Basin, Chile: reliability of water rights and changes in the demand for
irrigation. J Water Res Plan Manag 138(5):421–430
46. Calderón R, Palma P, Arancibia-Miranda N, Kim UJ, Silva-Moreno E, Kannan K (2020)
Occurrence, distribution and dynamics of perchlorate in soil, water, fertilisers, vegetables and
fruits and associated human exposure in Chile. Environ Geochem Health:1–9
47. Bonilla CA, Reyes JL, Magri A (2010) Water erosion prediction using the Revised Universal
Soil Loss Equation (RUSLE) in a GIS framework, central Chile. Chile J Agr Res 70(1):
159–169
48. Casanova M, Salazar O, Seguel O, Nájera F, Villarroel R, Leiva C (2012) Long-term
monitoring of soil fertility for agroforestry combined with water harvesting in Central Chile.
Arch Agron Soil Sci 58(sup1):S165–S169
49. Salazar O, Casanova M, Kätterer T (2011) The impact of agroforestry combined with water
harvesting on soil carbon and nitrogen stocks in central Chile evaluated using the ICBM/N
model. Agric Ecosyst Environ 140(1–2):123–136
50. Poblete D, Vicuña S, Meza F, Bustos E (2012) Water resources modelling under climate
change scenarios of Maule River Basin (Chile) with two main water intensive and competing
sectors: agriculture and hydropower generation. In: IWA world congress on water, climate and
energy
51. Salgado MMDPM (2012) Influence of compost and humic substances on soil and fruit quality
in table grape under intensive management in Chile. Doctoral Thesis Dissertation. University
of Bonn, Bonn, 138 p
52. Jara-Rojas R, Bravo-Ureta BE, Engler A, Díaz J (2013) An analysis of the joint adoption of
water conservation and soil conservation in Central Chile. Land Use Policy 32:292–301
53. Dörner J, Huertas J, Cuevas JG, Leiva C, Paulino L, Arumí JL (2015) Water content dynamics
in a volcanic ash soil slope in southern Chile. J Plant Nutr Soil Sci 178(4):693–702
54. Bown HE, Fuentes JP, Martínez AM (2018) Assessing water use and soil water balance of
planted native tree species under strong water limitations in Northern Chile. New For 49(6):
871–892
55. Gutiérrez-Gamboa G, Carrasco-Quiroz M, Verdugo-Vásquez N, Díaz-Gálvez I, Garde-
Cerdán T, Moreno-Simunovic Y (2018) Characterization of grape phenolic compounds of
‘Carignan’ grapevines grafted onto ‘País’ rootstock from Maule Valley (Chile): implications
of climate and soil conditions. Chil J Agri R 78(2):310–315
56. Lehnert LW, Thies B, Trachte K, Achilles S, Osses P, Baumann K, Schmidt J, Salomov E,
Jung P, Leinweber P, Karsten U, Büdel B, Bendix J (2018) A case study on fog/low stratus
occurrence at Las Lomitas, Atacama Desert (Chile) as a water source for biological soil crusts.
Aerosol Air Qual Res 18:254–269
57. Meza FJ, Montes C, Bravo-Martínez F, Serrano-Ortiz P, Kowalski AS (2018) Soil water
content effects on net ecosystem CO2 exchange and actual evapotranspiration in a Mediter-
ranean semiarid savanna of Central Chile. Sci Rep 8(1):1–11
58. Panez-Pinto A, Mansilla-Quiñones P, Moreira-Muñoz A (2018) Water, soil and
sociometabolic fracture of agribusiness. Fruit activity in Petorca, Chile. Bitác Urb Territ
28(3):153–160
59. Reyes Rojas LA, Adhikari K, Ventura SJ (2018) Projecting soil organic carbon distribution in
central Chile under future climate scenarios. J Environ Qual 47(4):735–745
Agriculture Land Degradation in Chile 147

60. Brucker E, Spohn M (2019) Formation of soil phosphorus fractions along a climate and
vegetation gradient in the Coastal Cordillera of Chile. Catena 180:203–211
61. Francke S, Carrasco L, Carnieletto C, Gándara E, Troncoso J (2019) Soil and water conser-
vation technics as a mechanism to adapt to the impacts of climate change in the Maule River
Basin Chile. In: Global symposium on soil erosion, Roma, 310 p
62. Castillo P, Serra I, Townley B, Aburto F, López S, Tapia J, Contreras M (2021) Biogeochem-
istry of plant essential mineral nutrients across rock, soil, water and fruits in vineyards of
Central Chile. Catena 196:104905
63. Fuentes I, Seguel O, Casanova M (2013) Elasto-plastic behaviour of soil aggregates and the
soil matrix as a function of physical properties in three soils of central Chile. Soil degradation.
Adv Geoecol 42:72–88
64. Seguel CG, Muñoz H, Segovia J, Ávalos B, Martín JR (2019) Assessment of soil contamina-
tion in caleta vitor and surrounding areas, northern Chile, due to heavy metal enrichment
caused by an abandoned copper mine. Interciencia 44(4):241–246
65. Hernández Á, Arellano EC, Morales-Moraga D, Miranda MD (2016) Understanding the effect
of three decades of land use change on soil quality and biomass productivity in a Mediterra-
nean landscape in Chile. Catena 140:195–204
66. Schuller P, Castillo A, Walling DE, Iroume A (2011) Use of fallout Caesium-137 and
Beryllium-7 to assess the effectiveness of changes in tillage systems in promoting soil
conservation and environmental protection on agricultural land in Chile. In: International
Atomic Energy Agency (ed) Impact of soil conservation measures on erosion control and
soil quality, pp 241–257
67. Fajardo A, Gundale MJ (2015) Combined effects of anthropogenic fires and land-use change
on soil properties and processes in Patagonia, Chile. Forest Ecol Manag 357:60–67
68. Fleige H, Beck-Broichsitter S, Dörner J, Goebel MO, Bachmann J, Horn R (2016) Land use
and soil development in southern Chile: effects on physical properties. J Soil Sci Plant Nutr
16(3):818–831
69. Henriquez-Dole LE, Vicuna S, Gironas JA, Meza FJ (2016) Adaptation measures evaluation
on agriculture under future climate and land use scenarios in central Chile. In AGU Fall
Meeting Abstracts H51A, p 1408
70. Dörner J, Horn R, Dec D, Wendroth O, Fleige H, Zúñiga F (2017) Land-use-dependent change
in the soil mechanical strength and resilience of a shallow volcanic ash soil in southern Chile.
Soil Sci Soc Am J 81(5):1064–1073
71. Soto L, Galleguillos M, Seguel O, Sotomayor B, Lara A (2019) Assessment of soil physical
properties’ statuses under different land covers within a landscape dominated by exotic
industrial tree plantations in south-central Chile. J Soil Water Conserv 74(1):12–23
72. Martínez IG, Prat C, Ovalle C, del Pozo A, Stolpe N, Zagal E (2012) Subsoiling improves
conservation tillage in cereal production of severely degraded Alfisols under Mediterranean
climate. Geoderma 189:10–17
73. Brunel-Saldias N, Seguel O, Ovalle C, Acevedo E, Martínez I (2018) Tillage effects on the soil
water balance and the use of water by oats and wheat in a Mediterranean climate. Soil Tillage
Res 184:68–77
74. Climent MJ, Herrero-Hernández E, Sánchez-Martín MJ, Rodríguez-Cruz MS, Pedreros P,
Urrutia R (2019) Residues of pesticides and some metabolites in dissolved and particulate
phase in surface stream water of Cachapoal River basin, central Chile. Environ Pollut 251:90–
101
75. Sheng P, Shang X, Sun Z, Yang L, Guo X, Jones MK (2018) North-south patterning of millet
agriculture on the Loess Plateau: late neolithic adaptations to water stress, NW China.
Holocene 28(10):1554–1563
76. Vila-Traver J, Aguilera E, Infante-Amate J, de Molina MG (2021) Climate change and
industrialisation as the main drivers of Spanish agriculture water stress. Sci Total Environ
760:143399
148 M. Francos

77. Sabbaghi MA, Nazari M, Araghinejad S, Soufizadeh S (2020) Economic impacts of climate
change on water resources and agriculture in Zayandehroud river basin in Iran. Agr Water
Manage 241:106323
78. Hendricks NP (2018) Potential benefits from innovations to reduce heat and water stress in
agriculture. J Assoc Environ Resour Econ 5(3):545–576
79. El-Areed SR (2019) Improvement of yellow corn productivity under water deficit stress using
conservation agriculture system. Egypt J Plant Breed 23(1):1–22
80. Adams RM, Rosenzweig C, Peart RM, Ritchie JT, McCarl BA, Glyer JD, Curry RB, Jones
JW, Boote KJ, Allen LH (1990) Global climate change and US agriculture. Nature 345(6272):
219–224
81. Agovino M, Casaccia M, Ciommi M, Ferrara M, Marchesano K (2019) Agriculture, climate
change and sustainability: the case of EU-28. Ecol Indic 105:525–543
82. Wegren SK (2021) Vulnerabilities in Russian agriculture to climate change. Sustain Dev Res
3(1):1–p1
83. Lehtonen HS, Aakkula J, Fronzek S, Helin J, Hildén M, Huttunen S, Kaljonen M, Niemi J,
Palosuo T, Pirttioja N, Rokkonen P, Varho V, Carter TR (2021) Shared socioeconomic
pathways for climate change research in Finland: co-developing extended SSP narratives for
agriculture. Reg Environ Chang 21(1):1–16
84. Cui X (2020) Climate change and adaptation in agriculture: evidence from US cropping
patterns. J Environ Econ Manag 101:102306
85. Chen S, Gong B (2021) Response and adaptation of agriculture to climate change: evidence
from China. J Dev Econ 148:102557
86. Piao S, Ciais P, Huang Y, Shen Z, Peng S, Li J, Zhou L, Liu H, Ma Y, Ding Y,
Friedlingstein P, Liu C, Tan K, Yu Y, Zhang T, Fang J (2010) The impacts of climate change
on water resources and agriculture in China. Nature 467(7311):43–51
87. Meza FJ, Vicuña S, Jelinek M, Bustos E, Bonelli S (2014) Assessing water demands and
coverage sensitivity to climate change in the urban and rural sectors in central Chile. J Water
Clim Change 5(2):192–203
88. Melo O, Foster W (2021) Agricultural and forestry land and labor use under long-term climate
change in Chile. Atmos 12(3):305
89. Guibal R, Lissalde S, Leblanc J, Cleries K, Charriau A, Poulier G, Mazzella N, Rebillard JP,
Brizard Y, Guibaud G (2018) Two sampling strategies for an overview of pesticide contam-
ination in an agriculture-extensive headwater stream. Environ Sci Pollut R 25(15):
14280–14293
90. Xiang T, Malik TH, Nielsen K (2020) The impact of population pressure on global fertiliser
use intensity, 1970–2011: an analysis of policy-induced mediation. Technol Forecast Soc 152:
119895
91. Hedlund J, Longo SB, York R (2020) Agriculture, pesticide use, and economic development: a
global examination (1990–2014). Rural Sociol 85(2):519–544
92. Zhou XY, Wang X (2019) Cd contamination status and cost-benefits analysis in agriculture
soils of Yangtze River basin. Environ Pollut 254:112962
93. Mahmood-ul-Hassan M, Yousra M, Ahmad R, Sarwar S (2019) Arsenic contamination in rice
grown under anaerobic condition in arid agriculture: assessment and remediation. Bull Envi-
ron Contam Toxicol 103(6):865–870
94. Hussain Z, Alam M, Khan MA, Asif M, Shah MA, Khan S, Khan S, Nawab J (2020)
Bioaccumulation of potentially toxic elements in spinach grown on contaminated soils
amended with organic fertilisers and their subsequent human health risk. Arab J Geosci
13(18):1–9
95. Ning CHEN, Shuai W, Xinmei HAO, Zhang H, Dongmei ZHOU, Juan GAO (2017) Con-
tamination of phthalate esters in vegetable agriculture and human cumulative risk assessment.
Pedosphere 27(3):439–451
Agriculture Land Degradation in Chile 149

96. Niedobova J, Skalský M, Fric ZF, Hula V, Brtnický M (2019) Effects of so-called “environ-
mentally friendly” agrochemicals on the harlequin ladybird Harmonia axyridis (Coleoptera:
Coccinelidae). Eur J Entomol 116:173–177
97. Sharma K, Cheng Z, Grewal PS (2015) Relationship between soil heavy metal contamination
and soil food web health in vacant lots slated for urban agriculture in two post-industrial cities.
Urban Ecosyst 18(3):835–855
98. Alekseev I, Abakumov E (2020) 13C-NMR spectroscopy of humic substances isolated from
the agricultural soils of Puchuncavi (El Melón and Puchuncavi areas), central Chile. Soil
Water Res 15(3):191–198
99. Vega M, Nerenberg R, Vargas IT (2018) Perchlorate contamination in Chile: legacy, chal-
lenges, and potential solutions. Environ Res 164:316–326
100. Calderón R, Rajendiran K, Kim UJ, Palma P, Arancibia-Miranda N, Silva-Moreno E,
Corradini F (2020) Sources and fates of perchlorate in soils in Chile: a case study of
perchlorate dynamics in soil-crop systems using lettuce (Lactuca sativa) fields. Environ Pollut
264:114682
101. Reyes A, Thiombane M, Panico A, Daniele L, Lima A, Di Bonito M, De Vivo B (2020)
Source patterns of potentially toxic elements (PTEs) and mining activity contamination level
in soils of Taltal city (northern Chile). Environ Geochem Health 42(8):2573–2594
102. Reyes A, Cuevas J, Fuentes B, Fernández E, Arce W, Guerrero M, Letelier MV (2021)
Distribution of potentially toxic elements in soils surrounding abandoned mining waste
located in Taltal, Northern Chile. J Geochem Explor 220:106653
103. Verdejo J, Ginocchio R, Sauvé S, Salgado E, Neaman A (2015) Thresholds of copper
phytotoxicity in field-collected agricultural soils exposed to copper mining activities in
Chile. Ecotox Environ Safe 122:171–177
104. Moya H, Verdejo J, Yáñez C, Álvaro JE, Sauvé S, Neaman A (2017) Nitrification and nitrogen
mineralization in agricultural soils contaminated by copper mining activities in Central Chile. J
Soil Sci Plant Nutr 17(1):205–213
105. Stowhas T, Verdejo J, Yáñez C, Celis-Diez JL, Martínez CE, Neaman A (2018) Zinc alleviates
copper toxicity to symbiotic nitrogen fixation in agricultural soil affected by copper mining in
central Chile. Chemosphere 209:960–963
106. Pancetti F, Ramírez M, Castillo C (2011) Epidemiological studies of anticholinestarase
pesticides poisoning in Chile. In: Satoh T, Gupta RC (eds) Anticholinesterase pesticides:
metabolism, neurotoxicity, and epidemiology. Wiley, Hoboken, pp 357–364
107. Ramírez-Santana M, Zúñiga-Venegas L, Corral S, Roeleveld N, Groenewoud H, Van der
Velden K, Scheepers PTJ, Pancetti F (2020) Association between cholinesterase’s inhibition
and cognitive impairment: a basis for prevention policies of environmental pollution by
organophosphate and carbamate pesticides in Chile. Environ Res 186:109539
108. Pancetti F, Olmos C, Dagnino-Subiabre A, Rozas C, Morales B (2007) Noncholinesterase
effects induced by organophosphate pesticides and their relationship to cognitive processes:
implication for the action of acylpeptide hydrolase. J Toxicol Environ Health B Crit Rev 10(8):
623–630
109. Dike S, Apte S (2020) A bibliometric analysis of soil pollution due to microplastics. Libr Phil
Pract:1–24
110. da Costa JP, Paço A, Santos PS, Duarte AC, Rocha-Santos T (2019) Microplastics in soils:
assessment, analytics and risks. Environ Chem 16(1):18–30
111. Zhou B, Wang J, Zhang H, Shi H, Fei Y, Huang S, Tong Y, Wen D, Luo Y, Barceló D (2020)
Microplastics in agricultural soils on the coastal plain of Hangzhou Bay, east China: multiple
sources other than plastic mulching film. J Hazard Mater 388:121814
112. Zhang S, Wang J, Yan P, Hao X, Xu B, Wang W, Aurangzeib M (2020) Non-biodegradable
microplastics in soils: a brief review and challenge. J Hazard Mater 409:124525
113. Corradini F, Meza P, Eguiluz R, Casado F, Huerta-Lwanga E, Geissen V (2019) Evidence of
microplastic accumulation in agricultural soils from sewage sludge disposal. Sci Total Environ
671:411–420
150 M. Francos

114. Allen S, Allen D, Phoenix VR, Le Roux G, Jiménez PD, Simonneau A, Binet S, Galop D
(2019) Atmospheric transport and deposition of microplastics in a remote mountain catch-
ment. Nat Geosci 12(5):339–344
115. Zhu F, Zhu C, Wang C, Gu C (2019) Occurrence and ecological impacts of microplastics in
soil systems: a review. Bull Environ Contam Toxicol 102(6):741–749
116. Klingelhöfer D, Braun M, Quarcoo D, Brüggmann D, Groneberg DA (2020) Research
landscape of a global environmental challenge: microplastics. Water Res 170:115358
117. Pazienza P, De Lucia C (2020) For a new plastics economy in agriculture: policy reflections on
the EU strategy from a local perspective. J Clean Prod 253:119844
118. Li J, Guo K, Cao Y, Wang S, Song Y, Zhang H (2021) Enhance in mobility of oxytetracycline
in a sandy loamy soil caused by the presence of microplastics. Environ Pollut 269:116151
119. Zhang S, Wang J, Liu X, Qu F, Wang X, Wang X, Li Y, Sun Y (2019) Microplastics in the
environment: a review of analytical methods, distribution, and biological effects. Trends Anal
Chem 111:62–72
120. Ng EL, Lwanga EH, Eldridge SM, Johnston P, Hu HW, Geissen V, Chen D (2018) An
overview of microplastic and nanoplastic pollution in agroecosystems. Sci Total Environ 627:
1377–1388
121. Ariza-Tarazona MC, Villarreal-Chiu JF, Hernández-López JM, De la Rosa JR, Barbieri V,
Siligardi C, Cedillo-González EI (2020) Microplastic pollution reduction by a carbon and
nitrogen-doped TiO2: effect of pH and temperature in the photocatalytic degradation process. J
Hazard Mater 395:122632
122. Napper IE, Thompson RC (2019) Environmental deterioration of biodegradable,
oxo-biodegradable, compostable, and conventional plastic carrier bags in the sea, soil, and
open-air over a 3-year period. Environ Sci Technol 53(9):4775–4783
123. Okpoho NA (2018) Smallholder agriculture land use impact on soil organic carbon stock in
federal capital territory of Nigeria. J Agric Environ Inte Develop 112(1):109–119
124. Devátý J, Dostál T, Hösl R, Krása J, Strauss P (2019) Effects of historical land use and land
pattern changes on soil erosion – case studies from Lower Austria and Central Bohemia. Land
Use Policy 82:674–685
125. Ferreira V, Panagopoulos T, Cakula A, Andrade R, Arvela A (2015) Predicting soil erosion
after land use changes for irrigating agriculture in a large reservoir of southern Portugal. Agric
Agric Sci Proc 4:40–49
126. León-Muñoz J, Echeverría C, Marcé R, Riss W, Sherman B, Iriarte JL (2013) The combined
impact of land use change and aquaculture on sediment and water quality in oligotrophic Lake
Rupanco (North Patagonia, Chile, 40.8 S). J Environ Manage 128:283–291
127. Rosero DC (2015) Análisis de la dinámica del cambio de uso del suelo en la Provincia de
Osorno, Región de los Lagos, Chile, periodo 1998-2013. Tesis de Magister, Universidad
Austral de Chile, Valdivia, 46 p
128. Montoya-Tangarife C, De La Barrera F, Salazar A, Inostroza L (2017) Monitoring the effects
of land cover change on the supply of ecosystem services in an urban region: a study of
Santiago-Valparaíso, Chile. PLoS One 12(11):e0188117
129. Nahuelhual L, Carmona A, Lara A, Echeverría C, González ME (2012) Land-cover change to
forest plantations: proximate causes and implications for the landscape in south-central Chile.
Landscape Urban Plan 107(1):12–20
130. Rodríguez-Echeverry J, Echeverría C, Oyarzún C, Morales L (2018) Impact of land-use
change on biodiversity and ecosystem services in the Chilean temperate forests. Landsc
Ecol 33(3):439–453
131. Braun AC, Banfield CC, Vogt J, Barra R, Schuller P, Koch B (2014) Assessing erosion risks
induced by land-use change in favor of commercial forestry in Chile. EARSeL eProceedings,
special issue: 34th EARSeL symposium, Warsaw, pp 1–9
132. Díaz GI, Nahuelhual L, Echeverría C, Marín S (2011) Drivers of land abandonment in
Southern Chile and implications for landscape planning. Landscape Urban Plan 99(3–4):
207–217
Agriculture Land Degradation in Chile 151

133. Carevic FS, Barrientos E, Anderson M (2017) Bodefales en el norte de Chile: una visión
general desde la perspectiva de los rasgos hidráulicos de la vegetación a la conservación
biológica. Idesia (Arica) 35(3):109–114
134. Sharma H, Shukla MK, Bosland PW, Steiner R (2017) Soil moisture sensor calibration, actual
evapotranspiration, and crop coefficients for drip irrigated greenhouse Chile peppers. Agr
Water Manage 179:81–91
135. Alcívar M, Zurita-Silva A, Sandoval M, Muñoz C, Schoebitz M (2018) Reclamation of saline-
sodic soils with combined amendments: impact on quinoa performance and biological soil
quality. Sustainability 10(9):3083
136. Parraguez-Vergara E, Contreras B, Clavijo N, Villegas V, Paucar N, Ther F (2018) Does
indigenous and campesino traditional agriculture have anything to contribute to food sover-
eignty in Latin America? Evidence from Chile, Peru, Ecuador, Colombia, Guatemala and
Mexico. Int J Agr Systain 16(4–5):326–341
137. Fierro P, Valdovinos C, Lara C, Saldías GS (2021) Influence of intensive agriculture on
benthic macroinvertebrate assemblages and water quality in the Aconcagua River Basin
(Central Chile). Water 13(4):492
Agricultural Land Degradation in China

Yang Yu, PanPan Ma, Qilin Zuo, Ming Gong, Miao Hu, and Paulo Pereira

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
2 Soil Compaction and Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
2.1 Effect of Soil Compaction on Soil Physicochemical Properties . . . . . . . . . . . . . . . . . . . . . 155
2.2 Soil Compaction Risk in Farmland Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
2.3 Technical Strategies for Avoiding or Mitigating Soil Compaction . . . . . . . . . . . . . . . . . . 157
2.4 Prospects of Research on Alleviating Soil Compaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
3 Overgrazing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
4 Slash and Burn Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
4.1 Classification of Slash-and-Burn Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
4.2 Solutions for Slash-and-Burn Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
4.3 Expectation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
5 Salinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
6 Agrochemical Use and Soil Contamination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
6.1 Current Hazard . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
6.2 Remediation of Farmland Soil Contaminated by Agricultural Chemicals . . . . . . . . . . 167
6.3 Comprehensive Control Countermeasures of Farmland Polluted by Agricultural
Chemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
6.4 The Way Forward . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
7 Microplastics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
7.1 Classification of Microplastics in Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
7.2 Sources of Microplastics in Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
7.3 Solutions to Microplastic Environmental Pollution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

Y. Yu (✉), Q. Zuo, and M. Gong


School of Soil and Water Conservation, Beijing Forestry University, Beijing, China
Jixian National Forest Ecosystem Research Network Station, CNERN, Beijing Forestry
University, Beijing, China
P. Ma
College of Agriculture and Animal Husbandry, Qinghai University, Xining, China
M. Hu
School of Landscape Architecture, Beijing Forestry University, Beijing, China
P. Pereira
Environmental Management Laboratory, Mykolas Romeris University, Vilnius, Lithuania

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 153
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 153–176, DOI 10.1007/698_2022_930,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 25 March 2023
154 Y. Yu et al.

7.4 Future Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170


8 Conclusion Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

Abstract China is a major agricultural country in the world. Since the 1980s,
China’s food problem has been basically solved, and 22% of the world’s population
has been fed and nursed with 9% of the world’s arable land. However, due to the
large population with relatively little land, the per capita arable land area is only
0.08 hm2, less than one-fourth of the world average. The hilly areas account for
about two-thirds of the total land area, and the middle and low yield fields account
for two-thirds of the total arable land area. There is a severe shortage of water
resources. China’s per capita water resources are only about 2,200 m3, listed as one
of the 13 water-deficient countries in the world. In recent years, the environment and
cultivated land quality have attracted government attention and related research.
Several policies were created to promote farmland system protection, agricultural
production, quality of agricultural products, and land degradation prevention and
control. However, what is the environmental condition of agricultural-producing
areas in China? How does this kind of condition influence the quality and safety of
agricultural products? This chapter summarises the main factors affecting China’s
farmland degradation from six aspects: soil compaction and erosion, overgrazing,
slash-and-burn agriculture, salinity, agrochemical use and soil contamination, and
microplastics.

Keywords Agricultural land degradation, Agrochemical use and soil


contamination, Microplastics, Overgrazing, Salinity, Slash-and-burn agriculture,
Soil compaction and erosion

1 Introduction

Soil, water, and biodiversity are the foundations of our societies and economies. A
global pattern of human dominance has resulted in persistent degradation and loss of
land resources in recent decades [1]. Degradation of soil results from natural or
anthropogenic disturbances that reduce actual or potential productivity or utility.
Agricultural land degradation is a global threat, and it has a strong influence on food
and energy resources and environments [2]. The degradation of soil, the quality of
the environment, food security, and energy consumption are all interconnected.
Globally, agricultural soil degradation is most often caused by human activities
and climate change. Agricultural soil degradation and accelerated soil erosion affect
agricultural productivity and the environment. Therefore, unveiling the complex and
interactive relationship between agricultural land degradation and soil erosion will
be a critical step towards solving the problem of food security and the decline in soil
Agricultural Land Degradation in China 155

environment quality [3, 4]. China is one of the largest agricultural countries [5]. Set-
tled agriculture may have started in China as early as 10,000 years ago. China’s
arable land protection and utilisation work face many challenges. The degradation of
agricultural land has intensified, the organic matter of the black soil in the northeast
has declined, the red and yellow soil in the south is acidified, the soil in the north is
arid and saline, and the area of degraded arable land with obstacles accounts for 40%
[6, 7]. In order to break the vicious cycle of soil erosion, low crop yield, poverty, and
low resource agriculture, policy and technological interventions need to be
implemented. In crops, microbes, soils, and plants interact and respond to different
management practices. Agroecosystems have become one of the most inextricable
self-organising systems [4, 8]. In contrast, how are the agricultural producing areas
in China affected by the environment? What are these conditions’ impacts on the
quality and safety of agricultural products? The scientific understanding of these
problems is of great significance to the prevention and control of agricultural soil
degradation, the healthy production and supply of agricultural products, and the
sustainable development of agriculture and the rural economy. Understanding how
human activities affect agricultural land degradation will result in management
practices and strategies to restore the soil environment to improve its quality and
prevent erosion while potentially improving the yield.
This chapter summarises the main factors affecting China’s farmland degrada-
tion: soil compaction and erosion, overgrazing, slash-and-burn agriculture, salinity,
agrochemicals use and soil contamination, and microplastics. Based on these pre-
mises, this chapter’s objectives are to review the impacts of the six aspects of the
abovementioned agricultural land degradation in China and highlight important
management strategies for preventing agricultural land degradation.

2 Soil Compaction and Erosion

Tractor trafficking increases soil compaction [9], which is considered one of the
most severe forms of degradation [10]. Agricultural soil compaction is common in
all world climatic regions. It is manifested by increased mechanical resistance and
bulk density and decreased porosity and water infiltration, affecting crop yields and
vegetation recovery [11]. Chinese scholars began to pay attention to the soil com-
paction problem in the 1980s. So far, the research carried out is mainly focused on
the influence of compaction on soil physical and chemical properties, improving soil
conditions, impacts on the carbon cycle, and modelling soil compaction [e.g. 12, 13].

2.1 Effect of Soil Compaction on Soil Physicochemical


Properties

The effect of compaction on soil physical properties is visible by the increase of soil
bulk density and the decrease in soil porosity, aeration, soil water holding capacity,
156 Y. Yu et al.

and permeability. Decreasing soil aeration and porosity decreases water infiltration,
microbiological activity, and root penetration [14].
Soil compaction directly increases soil bulk density and decreases soil porosity.
Under the same site condition, the more serious the soil compaction is the more
obvious surface soil bulk density and porosity changes. The effects of compaction on
soil bulk density, porosity, water permeability, and other soil structural parameters
are significant. The effects on soil physical properties will also increase with the
increase of compaction times. The change in soil structure caused by compaction is
mainly concentrated in the 0–30 cm soil layer. Meanwhile, as the degree of com-
paction increases, the soil’s thermal conductivity and heat capacity will both
increase [15].
Soil water retention is closely related to soil permeability. Soil water retention is a
comprehensive reflection of soil aggregates, soil porosity status, and particle com-
position, and soil porosity status plays a decisive role in water infiltration. Compac-
tion directly changes the composition of soil particles, the size and number of
aggregates, and the size and distribution of soil porosity, which inevitably affects
soil water retention capacity and permeability [16, 17]. This increases the likelihood
of surface runoff and soil erosion.
Compaction will affect the effectiveness of soil nutrient absorption and change
the adsorption and fixation process of chemical elements and ions in the soil. N, P,
and K are essential growth elements for plants. After compaction, soil aeration is
reduced, and surface runoff is increased, which directly leads to the loss of gaseous
nitrogen and nitrogen compounds. The compacted soil is not conducive to the
extension of plant roots, and the roots become thinner and smaller, while the uptake
of P and K mainly depends on the root configuration. At the same time, the decrease
in soil aeration leads to the decrease of O2 content in the soil, and the nutrient
elements such as Fe, Mn, Zn, and Cu will be released through peroxidation reduction
[10, 18, 19].

2.2 Soil Compaction Risk in Farmland Systems

The mechanical compaction of farmland soil not only significantly affects the
development of crop root systems and crop growth but also increases the energy
demand for cultivation, affects water infiltration and movement, and even causes soil
erosion and pollution [20, 21]. Mechanical compaction of soil has adverse effects on
soil physics, chemistry, biology, and crop growth and development, and its harms
mainly include the following:
1. Due to compaction, soil porosity decreases, natural macropores decrease, soil
water holding capacity decreases, and permeability decreases significantly. In the
case of short-term heavy rainfall, surface runoff increases, erosion increases, the
dry period is prolonged, and soil compaction has become one of the leading
causes of soil and water loss [22]. When soil porosity is less than 10%, soil
Agricultural Land Degradation in China 157

macropores are reduced, and small pores are closed by water. In these circum-
stances, the air exchange, carbon mineralisation, and nitrification are reduced.
Denitrification is enhanced, and the ratio of soil carbon and nitrogen will be
increased, resulting in a decrease in soil quality [23].
2. Compaction significantly changed the soil structure. The soil tended to be
anaerobic, the biomass and activity of soil microorganisms decreased, and the
number of soil animals decreased [24].
3. Compaction increases soil strength and osmotic resistance, which makes it
difficult for crop roots to grow in compacted soil. Root ratio in the surface layer
and secondary roots increase significantly. Roots near compacted areas become
coarser, flat, and tortuous, and epidermal cells are twisted and radiated [25].
4. Due to the increase of soil mechanical resistance, the energy consumption of soil
tillage increases, and the tillage benefit decreases. In the most severe case, the
positive benefit of tillage is offset by the negative benefit of compaction [26].

2.3 Technical Strategies for Avoiding or Mitigating Soil


Compaction

2.3.1 Soil Tillage Management

1. Avoid machine farming with high soil moisture


When the soil moisture content in the soil layer of 0.075–0.15 m is close to the
field capacity, the compaction caused by machinery to the soil is the most serious.
Therefore, the mechanical operation should be avoided when the soil moisture
content is high [27]. Cultivation should be done when the soil is at its driest. If it is
necessary to operate, the axle load should be reduced as far as possible, and the
tire grounding area should be increased to reduce the compaction effect on
the soil.
2. Reduced tillage
Soil aggregates can maintain good aeration, water movement, and root growth.
Traditional cultivation is too frequent, and it is easy to disturb the soil, break soil
aggregates, and worsen soil structure. Less tillage could reduce the sensitivity of
the soil to compaction and reduce the number of operations and the number of
times the soil is rolled [28]. Simultaneously, it could increase the amount of crop
residue on the surface, capture rainwater, and prevent surface crust formation
(Fig. 1).

3. Add organic matter to the soil


Add crop stubble, organic fertiliser, mud carbon, and green fertiliser to the
soil. Promote soil agglomeration, form a good structure, increase soil organic
matter content, and improve soil compaction resistance [29].
4. Crop rotation
158 Y. Yu et al.

Fig. 1 Farmland tillage methods on the Loess Plateau a: Abandoned farmland; b: Less-tillage
farmland; c: No-tillage farmland; d: No-tillage with straw return; e: tillage with straw return; f:
tillage

Crop rotation, including cereal–forage rotation, can significantly reduce soil


compaction. For example, the thick roots of alfalfa and clover can eliminate the
compaction of the core soil layer to a certain extent [e.g. 30, 31].
5. Change tillage depth
Because of the shearing effect of machines and tools on the soil, the plough
bottom is easily formed by constant depth cultivation. According to the soil
water content at the time of cultivation, the compaction area can be reduced by
appropriately changing the depth of cultivation [32]. Field investigations have
shown that deep ploughing is highly effective when soil compaction severely
restricts root growth and reduces crop yields. The proper deep ploughing
method should be selected according to the position of the compaction area in
the soil layer. If the plough layer is compacted, deep ploughing can be chosen. If
the core soil is compacted, the core soil shovel or chisel plough should be
selected for deep loosening. Deep ploughing is best carried out in autumn when
the soil moisture content is suitable [33].

2.3.2 Technical Improvement Measures

Machine axle load and grounding pressure are the main reasons for soil compaction
[34]. It is necessary to minimise the axle load and grounding pressure in tractor
design. The specific measures are as follows:
1. Change two-wheel drive to four-wheel drive
More than 70% of the total load of a two-wheel drive tractor is distributed in
the rear axle. Therefore, the rear wheels hit the ground harder. If changed to four-
Agricultural Land Degradation in China 159

wheel drive, at rest, the front axle load accounts for 55%. The rear axle load
accounts for 45%. When walking, the axle load distribution is more uniform,
which reduces the axle load and the rolling of soil without increasing the total
weight [35].
2. Change four-wheel to six-wheel or eight-wheel
Increasing the number of tires and tire width increases the ground area, and the
tire’s internal pressure can be reduced [36]. This can significantly reduce the
tractor pressure on the surface
3. Change wheeled tractor to tracked tractor
The ground area of a tracked tractor is much larger than that of a wheeled
tractor. Therefore, the compaction effect on the soil when cultivated with a
tracked tractor is much less than that of a wheeled tractor [37].
4. Change multilayer tire to radiate tire
Improve the traditional multilayer tire to radiate layer tire. When the inner
pressure of the tire is reduced, the ground area of the radial tire increases, and the
left and right sides remain unchanged. Therefore, the radial tire does not increase
the track area, which benefits the ridge planting or crops with small row spacing.
The ground area of the ordinary multilayer tire will increase, and the concave
change occurs in the middle of the tire crown, which causes soil shear. With the
increase of load or tire pressure decrease, the direction-finding rolling area of
common multilayer tires is increased, which may cause harm to crops [38].

2.4 Prospects of Research on Alleviating Soil Compaction

Compaction not only changes the physical and chemical properties of soil and
significantly reduces soil productivity, but also the CO2 produced by soil respiration
cannot be released in time, which affects the expected growth of plants and soil
biological activities. With the expansion of human activities in recent years, the area
of soil compaction is becoming larger and larger. If we do not pay more attention,
soil quality and the environment will deteriorate rapidly [39].
Currently, studies on the effects of compaction on soil under different land use
patterns mainly focus on the changes in soil’s physical and chemical properties
[40]. Studies on the effects of compaction on the CO2 release from soil respiration
are still weak, but it is an important step that cannot be ignored in studying the global
carbon cycle. Second, some soil compaction is inevitable when conducting produc-
tion operations, but it does not mean that soil compaction can only have adverse
effects. We know, as long as it does not exceed the natural resilience of the
environment. Compaction limits the amount of CO2 released from the soil surface
and can slow down the greenhouse effect. Future studies should pay attention to the
extent and scope of the effects of different compaction degrees on soil respiration
under different land use patterns. Suppose the scope of reasonable compaction
degree can be determined. In that case, the adverse effects of mechanical compaction
on the environment can be reduced by injecting advanced technology and improving
160 Y. Yu et al.

mechanical equipment. Finally, the study on the effect of compaction on soil


respiration and the degree of correlation between the effects of compaction on soil
respiration should be considered comprehensively, which will be of great help to
reveal the whole process of soil respiration and develop soil science research.
Soil mechanical compaction is an unavoidable problem in the development of
agricultural mechanisation. Appropriate methods should be used to analyse and
evaluate the characteristics of current agricultural machinery tillage compaction
and accurately describe the tire-soil compaction mechanism [36]. From the view-
point of environment protection, evaluate and select the structure types and param-
eters of walking devices, improve the walking device of existing machines, and
design new walking systems to reduce soil compaction and achieve sustainable
agricultural development.

3 Overgrazing

Overgrazing mainly occurs in China’s grassland ecosystem, especially in the agro-


pastoral zone. China is the second largest grassland country in the world. However,
due to the excessive exploitation and utilisation of grassland resources, the degra-
dation of the agro-pastoral region is severe. According to statistics, at present, 90%
of the grassland in China has been or is degrading. Among the 20,000 hm2 of
desertification lands that expand yearly, grassland is the main area affected [41].
Various factors lead to grassland degradation, including natural factors (e.g. long-
term drought, wind erosion, water erosion, sandstorms, rodents, insect pests), but
mainly human factors (e.g. overgrazing, over reclamation, woodcutting, mining).
Overgrazing is considered the leading cause of degradation (Fig. 2) in 30.1% of the
grasslands in China [42]. The influence of overgrazing on the grassland ecosystem
can be summarised as the influence on grass (plants) and the influence on land (soil).
Overgrazing not only negatively affects plant growth, development, and reproduc-
tive capacity but also changes the physical properties of the soil, as it results in
reduced cover and compaction of the soil due to trampling by livestock [43]. The
effects on plant communities and soil are considered destructive. In turn, these
processes increase soil crusting, reduce soil permeability, increase soil surface
erosion, reduce plant availability to soil water, and affect plant flora composition.
For local people, grazing is one of their primary sources of income. For this
reason, overgrazing has been a significant cause of grassland degradation [41]. The
decrease in available grassland area is the direct cause of overgrazing. On Inner
Mongolia grassland, the large area of grazing land used in the nomadic era is
decreasing due to the expansion of the occupied area of agricultural land, the
occupation of grassland for mining mineral resources, the occupation of residential
construction and grassland division needed for settlement, and the occupation of
grazing land for road construction. Livestock is more concentrated on the more
limited grassland, resulting in a significant contradiction between grass and live-
stock, and the grassland is gradually desertifying. The excessive pursuit of livestock
Agricultural Land Degradation in China 161

Fig. 2 Overgrazing leads to grassland desertification

increase is the indirect cause of overgrazing [44]. Farmers are primarily engaged in
traditional herding. Cattle and sheep are their only source of income. As rational
“economic men”, farmers also aim to maximise their interests. Driven by a strong
desire to improve living standards, farmers only have to look for ways from
grassland. Increasing livestock production is almost the only option for farmers to
increase their ability to cope with changes in economic incomes, ensure that living
standards do not decline even in a lousy year, and meet their expenses [45].
In order to solve the problems caused by overgrazing, several aspects must be
considered. (1) Clarify the property rights of grasslands and end the excessive use of
grasslands institutionally. First of all, we should continue to implement “dual rights
and one system” so that herdsmen can organically combine the interests of raising
livestock with the maintenance of grassland. In addition, we should reasonably
extend the paid contract period for pastures in light of local conditions. (2) Reduce
the pressure on utilising grasslands and apply periodical rotations. To reduce the
pressure of grassland utilisation, we should reduce the pressure on the population in
pastoral areas and the number of livestock. (3) Strengthen the construction of
grasslands and improve the capacity of grass production and livestock carrying.
(4) Change the production mode and reduce animal husbandry industrialisation.
Seeking economic growth solely by increasing primary livestock products will
inevitably lead to the overuse of grassland. Therefore, the sustainable development
of grassland animal husbandry must be combined with the deep processing of animal
162 Y. Yu et al.

products and the international division of labour of animal products and implement
industrial management.

4 Slash and Burn Agriculture

As a traditional agricultural production mode, slash and burn plays an undoubted


role in developing and promoting ancient agriculture in China [46]. However, we
must see that slash-and-burn and agricultural development promoted by it will affect
the environment and even expand the impact on the environment. Slash-and-burn
farming has always been an important agricultural production mode in ancient
society, and it was also the traditional means of acquiring farmland in various
countries then. As a result, slash-and-burn farming lasts for a long time and has a
wide range. Such long-term slash-and-burn farming on a large area will affect the
environment and lead to ecological changes [47].
Slash-and-burn farming is a necessary economic and cultural type of agricultural
development. It has been gradually replaced with the improvement of agricultural
productivity. However, in the minority areas in southwest China, this farming
method still exists today and even prevails in some areas. For example, slash-and-
burn farming still exists in some minority areas in Yunnan province, especially in
remote mountainous areas. Among the 25 ethnic minorities living in Yunnan
province, more than two-thirds of them followed the fire-tillage method. At the
beginning of the twenty-first century, while modern agricultural science and tech-
nology were vigorously promoted, the method of fire-tillage was still being quietly
applied in some areas. In the total land area of Yunnan, mountains and plateau
account for about 94%, and basins and broad valleys account for 6%. The average
altitude is over 2,000 m. Plain dams and gentle slopes of less than 15 degrees in the
province only account for 22% of the total land area, 15–25 degrees for 38%, and
steep slopes over 25 degrees for 40%. The characteristics of the geographical
environment limit the choice of livelihood modes of various ethnic minorities in
Yunnan to a large extent. Some minority areas can only live and multiply by slash-
and-burn agriculture [48].

4.1 Classification of Slash-and-Burn Patterns

1. Slash and burn without rotation


This type of farming is characterised by the cultivation of a single season of crops
on a plot of land, also known locally as lazy fire land, which is slash and burn at a
relatively low level of productivity [46]. Almost all the ethnic groups in Yunnan
mainly practised slash-and-burn farming in the early period, and waste farming
was their unique farming system.
2. Rotational slash and burn
Agricultural Land Degradation in China 163

This type of farming is characterised by cultivating crops continuously on


reclaimed land for several years and then allowing the land to lie fallow for
several years until the land is restored. Unlike the former, this slash-and-burn
farming system is no longer pure “waste farming” but has the colour of a “fallow
rotation” system. Furthermore, it is not simply ignored when the land is at leisure.
It is not only restored by the power of nature. For example, the Wa, Jingpo,
Dulong, Nu, and other nationalities plant the fast-growing species – water winter
gourd tree on the leisure land to cultivate good vegetation and shorten the leisure
period. The Hani people in Mengsong and Jinghong plant rattan and bamboo on
the leisure land and use them as bamboo fences. A few years later, the bamboo
will be cut, and the land will be burned to cultivate ginger and tobacco. It is an
excellent leap of minority slash-and-burn culture from restoring the ecology by
the force of nature to adjusting the relationship between nature, population, and
economy by the active role of humans [49].

4.2 Solutions for Slash-and-Burn Agriculture

4.2.1 Exploration of Alternatives to Slash-and-Burn Agriculture

In the 1980s, the land use in the Tiaoheba area underwent significant changes, and
the agricultural economic structure began to adjust. The primary feature was the
change from rice-based food agriculture to a new type of agriculture with the
development of cash crops. Based on self-sufficiency in grain, they found and
formed the primary road of economic development, planting amomum and rubber
and developing aquaculture. Since the 1990s, they have practised intensive farming.
Despite modern farming techniques, they do not use any fertilisers on cultivated
land. Instead, they focus on managing the farmland, the irrigation system, and the
surrounding environment. The agricultural economic structure gradually improved,
and farmers’ income gradually increased. In the twenty-first century, the growth of
economic forests in Xishuangbanna, especially the widespread planting of rubber
forests, has indeed increased farmers’ income in some ways, and most farmers have
entirely gotten rid of the slash-and-burn mode of production. However, on the other
hand, there are still some hidden dangers in environmental protection. Planting a
large area of rubber forest is not conducive to protecting the environment in the long
run [50].

4.3 Expectation

As one type of agriculture, slash and burn is the adaptation of tropical and subtrop-
ical mountain peoples to their habitats. Slash-and-burn practices are traditional in
mountain areas, and however, they can also pose environmental problems
164 Y. Yu et al.

[49]. Slash and burn is ostensibly a type of agricultural production but is a profound
social problem, especially in ethnic minority areas in Yunnan. It involves specific
areas, ethnic groups, environmental conditions, social systems, culture, customs,
politics, religion, and other related aspects. The slash-and-burn practices applied by
Yunnan minorities should be fully explored, such as the traditional rules of forest
protection, rational utilisation of natural resources, and the idea of harmonious
coexistence between man and nature. This is key to preserving the culture
and religion of Yunnan ethnic minorities [51]. We should choose the right attitude
and methods to treat slash and burn, take measures according to local conditions, and
strive to improve economic and cultural development and people’s living standards
in minority areas.

5 Salinity

Salinised soil is widely distributed in China. However, it is more common in the


Northern and coastal areas (Fig. 3). A large amount of salinised soil is also distrib-
uted on cultivated land. Saline soil is one of China’s most important causes of
low-yielding soil types, and human activities significantly affect it [52]. In recent
years, with the increasing use of exploitable land, people have turned their attention
to arid and semi-arid areas. However, due to the lack of water resources in these
areas, agricultural water saving has become an important issue in water resources
management. Moreover, due to the alternating effect of rainfall, irrigation, and
evaporation, the salt accumulates in the unsaturated soil, resulting in secondary
salinisation. Soil salinisation has become a global environmental problem faced by
the whole world [53]. Soil salinisation is related to resources and the environment,
and it directly affects food production and is a major restriction and obstacle to
agricultural development and sustainable development. Therefore, traditional and
new methods, such as remote sensing, are key to identifying and mitigating their
impacts.

Fig. 3 Salinised soil formed by abandoned farmland in China


Agricultural Land Degradation in China 165

The process and mechanism of soil salinisation are complicated. Its formation is a
form of land degradation influenced by natural factors and human activities. The
fragile eco-geological environment in arid and semi-arid areas is the objective basis
of soil salinisation. Geological conditions, soil conditions, low-level terrain, high
wind, groundwater, groundwater salinity, evaporation, salt-containing parent mate-
rial, drought, and less rain climate conditions are important natural factors that affect
soil salinisation [53]. Also, unsustainable human practices such as overexploitation,
overgrazing, and irrigation accelerated soil salinisation. As an important type of soil
degradation, the relationship between soil salinisation and the environment has been
paid more and more attention. Recently scientific and technical works have been
carried out to assess the human and climate impacts on soil salinisation [e.g. 54].
Human and natural factors can be responsible for the soil salinisation process
together with natural factors. Tillage and excessive irrigation are some of the causes
responsible for secondary salinity [55]. Human factors have two effects on
salinisation risk. The first is a positive effect; if human development and utilisation
and irrigation methods are unsustainable, the salinisation risk will increase. The
second is a reverse effect [56].
1. Unsustainable human production and living activities have promoted the occur-
rence of soil salinisation. In the past, people unilaterally implemented the policy
of focusing on grain and blindly destroying grass to open up land, which led to the
reclamation of a considerable part of grassland suitable for grazing into arable
land. After grassland was reclaimed, the surface “desalinated layer” as a result of
the growth of vegetation was destroyed. Salt rose to the surface with evaporation
from the deep level, resulting in salinity [57].
2. Excessive agricultural water use and backward irrigation technology have inten-
sified the frequency and intensity of the conversion between groundwater and
surface water, resulting in significant changes in the chemical characteristics of
regional groundwater and surface water [58]. At the same time, the groundwater
level of some wasteland or half wasteland in and around the irrigated area will rise
and become a dry wasteland, leading to the accumulation of soil salt in these areas
and becoming a hidden danger of soil salinisation.
3. Humans can influence the salinisation by building drainage channels. Ground-
water level and the saline-alkali components of the soil itself play a significant
role. The alkaline drainage channel is proposed based on the above principles. It
effectively reduces the groundwater level, discharges the saline-alkali compo-
nents, and controls salinisation [59].
With further research on the essential characteristics of saline soil and its evalu-
ation techniques, the comprehensive application of multidisciplinary methods and
techniques has been strengthened. The optimal management of saline soil utilisation
and the optimal control of soil water and salt dynamics are emphasised [60]. The
influence of human activities on secondary soil salinisation has been gradually
linked with environmental change [61]. Environment-friendly and cost-saving bio-
logical treatments and salt-soil farming techniques are highlighted elsewhere [62]
The research on saline soils needs to clarify the transport mechanism of salts into the
166 Y. Yu et al.

soil and their relationship with the plants [61]. In saline soils, management tech-
niques should be utilised to optimise the regulation of the saline soil’s water, salt,
and fertility. Soil quality could be improved, and secondary salinisation could be
prevented by directional cultivation of saline soil to improve the productivity of
saline soil and the utilisation efficiency of saline soil resources in China [61].
In general, future research on soil salinisation needs to pay attention to six
aspects: (1) monitoring, assessment, prediction, and early warning of soil
salinisation; (2) study of soil water and salt transport process at field scale and its
simulation; (3) interaction mechanism between plant and soil salinity and biological
treatment of saline soil; (4) study on mechanism and technology of soil water and
salt optimisation; (5) study on the treatment and restoration of saline-alkali barrier
and optimal management of saline soil resource utilisation; and (6) environmental
effects of soil salinisation.

6 Agrochemical Use and Soil Contamination

As an indispensable means of production in agricultural production, agricultural


chemicals have made significant contributions to agricultural development and
human food supply. However, in recent years, with the long-term and massive
application of agricultural chemicals, the problems of residues and pollution of
agricultural chemicals have become increasingly severe and become one of the
important sources of agricultural non-point source pollution [63]. China is a tradi-
tional agricultural country, and the use of agricultural chemicals ranks first in the
world, reaching 500,000–600,000 tons per year, of which 80–90% will eventually
enter the soil environment, resulting in about 870,000–1.07 million hm2 of farmland
soil polluted by agricultural chemicals [64]. In China, Shanghai, Zhejiang, Shan-
dong, Jiangsu, and Guangdong are regions with the largest use of agricultural
chemicals, among which Shanghai and Zhejiang reach the highest amount of
10.8 kg/hm2 and 10.41 kg/hm2, respectively [65].

6.1 Current Hazard

The actual harm caused by chemicals is mainly reflected in the harm to the environ-
ment and human health. Firstly, from the perspective of the environment, it is mainly
reflected in three aspects: air pollution, water pollution, and soil pollution. Hazard-
ous chemicals in industrial waste gas and dust, automobile exhaust, and agricultural
chemicals have caused severe air pollution, mainly reflected in the north of China’s
air pollution [66]. Industrial organic products and agricultural chemicals cause
severe water pollution and water quality deterioration through the discharge and
leakage of wastewater. Harmful chemicals enter rivers, lakes, seas, and farmland,
causing water and severe soil pollution [67]. Also, soil acidification is a consequence
Agricultural Land Degradation in China 167

of acid rain. All these events hurt human health. Many harmful substances enter
people’s respiratory system, causing respiratory diseases, heart and lung diseases,
and even cancer. Water pollution causes more human diseases and even directly
causes death in severe cases. On the one hand, soil pollution causes destructive
effects on agricultural production. On the other hand, it causes great harm to human
health through agricultural products [68].

6.2 Remediation of Farmland Soil Contaminated by


Agricultural Chemicals

There are several methods to remediate agriculture soil pollution (e.g. physical,
chemical, and biological). Physical and chemical techniques are efficient but very
expensive and require much work. They are more suitable for remediating soils with
high concentration pollutants [69]. Bioremediation has a long repair cycle. It is more
suitable for soil contaminated by agricultural chemicals with medium and low
residual concentration, such as remediation of agrochemical pollution in farmland
soil, because of its advantages such as financial and environmental protection, and
not easy to damage the ecosystem. Soil bioremediation includes microbial remedi-
ation, phytoremediation, and mycorrhizal remediation [70].

6.3 Comprehensive Control Countermeasures of Farmland


Polluted by Agricultural Chemicals

The pollution of harmful chemicals to the environment is evident to all. The effective
prevention and control of pollution has become an urgent matter for the development
of industry and agriculture in China at the present stage [71]. Under the premise of
correct understanding of environmental hazards caused by harmful chemicals,
strengthening the investigation and research on the residues of agricultural chemicals
in farmland soil, increasing the research and development of alternative technologies
for harmful agricultural chemicals, adjusting agronomic measures and strengthening
the self-purification capacity of the soil, guiding peasants to rational drug use and
safe application technology, improving their awareness of environmental protection,
perfecting laws and regulations, and establishing a standard quality system in line
with international standards and other measures can all provide practical solutions
and means for comprehensive treatment of polluted farmland [67].

6.4 The Way Forward

China’s soil contamination has specific development characteristics, and the preven-
tion and control of soil pollution has become one of the most important aspects of
168 Y. Yu et al.

China’s agricultural and environmental protection. The pollution of the environment


by harmful chemicals is evident to all [72]. How to prevent and control pollution has
become an urgent matter for developing industry and agriculture in China. The
control of pollution sources is a practical way to avoid soil pollution. The application
of cleaner production technology provides a reliable guarantee for controlling
pollution sources. It is more important to adopt appropriate production technologies
or reasonable development plans to minimise soil pollution and its hazards.
Prioritising the development of ecological restoration technologies for contaminated
soil is one of the most urgent tasks in soil pollution prevention and control in the
future [73]. Future work in this field will focus on researching soil compound
pollution related to food safety early warning and soil environmental quality bench-
mark determination. The problem of agricultural non-point source pollution and the
problem of soil environmental pollution in livestock and poultry production need to
be taken seriously. In order to protect cultivated land and prevent soil pollution, the
protection and sustainable use of land has become an important agenda for land
management departments and government agencies at all levels. As an important
part of agricultural and environmental protection, it is necessary to establish a soil
environmental pollution monitoring network and the most stringent chemical and
pesticide use specifications for continuous soil environmental management.

7 Microplastics

Environmental microplastic pollution is becoming one of the most severe threats to


the entire surface ecosystem of the earth, which has been a serious concern for the
governments of many countries around the world and is widely studied by scientific
circles. In recent years, the pollution problem of microplastics has been paid more
and more attention by Chinese scholars [74, 75]. Especially the accumulation of
microplastics in the soil environment is considerable, and it is not easy to degrade.
Therefore, the impact of long-term microplastic residues on the soil ecosystem and
its ecological effects have attracted widespread attention. The amount of
microplastics deposited into the soil exceeds the ones released into the ocean
[76]. This shows that microplastic pollution in agriculture is severe.

7.1 Classification of Microplastics in Soil

Microplastics can be divided into primary and secondary according to size [77]. Pri-
mary microplastics refer to the plastic microspheres and nano-plastics made into
micron size during the production process and used as raw materials in industrial
detergents and cosmetics. Secondary microplastics are fragments produced by large
plastic products, such as household garbage and plastic film. The shape of
microplastics can be divided into fibres, beads, films, flowers, spheres, particles,
Agricultural Land Degradation in China 169

foam, and debris [78]. It is essential to classify the particle size of microplastics,
which affects their migration potential in soil. Microplastics with particle size
>150 mm are unlikely to be absorbed by most plants and soil organisms (except
mesozoa) and pose no risk to human health.

7.2 Sources of Microplastics in Soil

The sources of microplastics in soil include agricultural mulching, sludge


composting, and atmospheric deposition, among which the large-scale use of agri-
cultural mulch is the most important factor. China is one of the major countries that
use plastic mulching films. Sludge also contains a large number of synthetic fibre
microplastics, up to 4.2–15.4 × 103 particles per kilogram of dry sludge
[79]. According to the total sludge production in China, it is estimated that
1.56 × 1014 a-1 microplastics will enter the soil through sludge [80]. The deposition
flux of microplastics in the atmosphere is up to 1.46 × 105 m-2 a-1. The fibre type
accounts for 95% [81, 82].

7.3 Solutions to Microplastic Environmental Pollution

1. Physical and chemical fragmentation and biodegradation of microplastics


After entering the environment, plastics crack and degrade into microplastics
[82]. As a powerful decomposer in nature, microorganisms can degrade almost all
known natural compounds and most synthetic substances. Plastic-degrading
microorganisms in terrestrial environments contain more than 30 genera of
bacteria and fungi, including Bacillus, Staphylococcus, Pseudomonas,
Achromatium, Amycolatopsis, Rhodococcus, filamentous fungi (Aspergillus),
and yeast [83]. Its separation sources include soil, landfills, compost, and animal
intestines [84]. The main degradable plastic polymers are polyethene, polyethene
terephthalate, polystyrene, polyvinyl chloride, polypropylene, and polyurethane.
Bacterial, fungal strains, microflora, and biofilms can degrade microplastics in the
environment [85]. Previous works showed that microorganisms degrade
microplastics by secreting extracellular enzymes such as hydrolase, esterase,
lipase, proteinase K, keratinase, urease, and oxidase [86].
2. Risk reduction strategies and technologies for microplastics
Source prevention is the best choice to control and reduce the environmental
risks of microplastics [82]. Microplastics are small in volume (nm~mm), and it
is almost difficult to recycle once they enter the environment. Therefore, it is
important to reduce plastic sources and eliminate the potential risks of
microplastic pollution [87]. Specifically speaking, in terms of the legal system,
we should build, revise, and improve relevant laws and regulations to provide
170 Y. Yu et al.

the legal basis for preventing and controlling plastic pollution. The economic
policy encourages waste plastics recycling and sets materials and product
standards to improve the quality of recycled plastic. Reducing plastic use at
source by limiting production, banning the use, product substitution and tax
regulation, and exploring the use of an extended producer responsibility system
to promote plastic waste recycling and reduce plastic waste in the environment
[88]. In terms of technological innovation, product design should be improved
to extend the service cycle of plastic products, and environmentally friendly
alternative materials should be researched and developed to reduce the use of
plastic raw materials [89]. It is essential to understand the ecological effects of
microplastics, assess their ecological risks and human health hazards, and
provide direct evidence of the harm of plastics. Regarding education, improve
public consciousness, change consumption patterns, develop citizen science,
and advocate public participation in preventing and controlling environmental
plastic waste (microplastics) pollution.

7.4 Future Development

Microplastics are widely distributed and accumulated in the soil, which has attracted
researchers’ attention to their source, distribution, migration, and ecological risk
[78]. However, due to the imperfection of appropriate research methods and classi-
fication standards, there is a lack of consistency in the research on the environmental
behaviour of microplastics in soil [77]. The research on the ecological risks of
microplastics to soil structure and physical and chemical properties, soil plants,
animals, and microorganisms is still in the exploration and accumulation stage and
has not yet formed a complete theoretical and methodological system [82]. In the
future, in-depth research should be carried out from the following aspects:
1. Establish unified microplastic sample collection, classification, separation, and
detection methods. Currently, the separation and detection methods of
microplastics in soil are limited. Therefore, it is urgent to standardise the separa-
tion and detection system of microplastics in soil.
2. Microplastics can adsorb heavy metals, organic pollutants, and antibiotics in soil.
Further research on the combined pollution of microplastics is needed, and it is
important to understand microplastic migration, biotransformation, and
bioaccumulation in the soil.
3. Research on ecological risk of microplastics to soil organisms. Future research
should include more microplastics, animal and plant species, and soil conditions.
The effects of microplastics and related pollutants should be further studied. It is
essential to understand that the interaction between microplastics and soil micro-
organisms, including microbial community structure and activity, should be
further studied. More field experiments should be carried out to study further
the effects of microplastic exposure on soil fauna under real-world environmental
scenarios.
Agricultural Land Degradation in China 171

8 Conclusion Remarks

In China, agricultural land degradation was focused mainly on foods required to


meet the basic demands for survival. Therefore, the current status and some specific
aspects must be taken seriously. In this chapter, we started from six aspects and
elaborated on the main factors that potentially affect agricultural land degradation in
China. Soil compaction, erosion, overgrazing, slash-and-burn salinity, agrochemical
use, and microplastics are the main causes of soil agricultural degradation in China.
The scientific understanding of these problems is of great significance to the
prevention and control of farmland soil degradation, the healthy production and
supply of agricultural products, and the sustainable development of agriculture and
the rural economy.

References

1. United Nations Convention to Combat Desertification (2022) The Global Land Outlook, 2nd
edn. UNCCD, Bonn
2. Alexandratos N, Bruinsma J. World agriculture: towards 2030/2050: the 2012 revision.
FAO, Rome
3. Pereira P, Barcelo D, Panagos P (2020) Soil and water threats in a changing environment.
Environ Res 186:109501
4. Viana C, Freire D, Abrantes P, Rocha J, Pereira P (2022) Agricultural land systems and related
research fields supporting food security and the Sustainable Development Goals: evidence from
a systematic review. Sci Total Environ 806:150718
5. Jia L, Zhao W, Zhai R, An Y, Pereira P (2020) Quantifying the effects of contour tillage in
controlling water erosion in China: a meta-analysis. Catena 195:104829
6. Du Z, Gao B, Ou C, Du Z, Yang J, Bayatungalag B, Battogtokh D, Yun W, Zhu D (2021) A
quantitative analysis of factors influencing organic matter concentration in the topsoil of black
soil in Northeast China based on spatial heterogeneous patterns. ISPRS Int J Geo-Inf 10:348
7. Han Y, Yi D, Ye Y, Liu S (2022) Response of spatiotemporal variability in soil pH and
associated influencing factors to land use change in a red soil hilly region in southern China.
CATENA 212:106074
8. Pereira P, Bogunovic I, Munoz-Rojas M, Brevik E (2018) Soil ecosystem services, sustainabil-
ity, valuation and management. Curr Opin Environ Sci Health 5:7–13
9. Alakukku L, Weisskopf P, Chamen WCT, Tijink FGJ, van der Linden JP, Pires S, Sommer C,
Spoor G (2003) Prevention strategies for field traffic-induced subsoil compaction: a review. Part
1. Machine/soil interactions. Soil Tillage Res 73:145–160
10. Nawaz MF, Bourrié G, Trolard F (2013) Soil compaction impact and modelling. A review.
Agron Sustain Dev 33:291–309
11. Głab T (2014) Effect of soil compaction and N fertilisation on soil pore characteristics and
physical quality of sandy loam soil under red clover/grass sward. Soil Tillage Res 144:8–19
12. Zhang XY, Cruse RM, Sui YY, Jhao Z (2006) Soil compaction induced by small tractor traffic
in Northeast China. Soil Sci Soc Am J 70:613–619
13. Wei B, Li Z, Wang Y (2022) Study on soil compaction and its causative factors at apple
orchards in the Weibei Dry Highland of China. Soil Use Manag 38:790–801
14. Hamza MA, Anderson WK (2005) Soil compaction in cropping systems: a review of the nature,
causes and possible solutions. Soil Tillage Res 82:121–145
172 Y. Yu et al.

15. Labelle ER, Jaeger D (2011) Soil compaction caused by cut-to-length forest operations and
possible short-term natural rehabilitation of soil density. For Range Wildland Soils 75:2314–
2329
16. Vomocil JA, Flocker WJ (1961) Effect of soil compaction on storage and movement of soil air
and water. Trans ASAE 4:0242–0246
17. Yu Y, Zhao WW, Martinez-Murillo JF, Pereira P (2020) Loess Plateau: from degradation to
restoration. Sci Total Environ 738:140206
18. Arvidsson J (1999) Nutrient uptake and growth of barley as affected by soil compaction. Plant
Soil 208:9–19
19. Kristoffersen A, Riley H (2005) Effects of soil compaction and moisture regime on the root and
shoot growth and phosphorus uptake of barley plants growing on soils with varying phosphorus
status. Nutr Cycling Agroecosyst 72(2):135–146
20. Bogunovic I, Pereira P, Kisic I, Sajko K, Sraka M (2018) Tillage management impacts on soil
compaction, erosion and crop yield in Stagnosols (Croatia). Catena 160:376–384
21. Bogunovic I, Pereira P, Galic M, Bilandzija D, Kisic I (2020) Tillage system and farmyard
impact on soil physical properties, CO2 emissions and crop yield in an organic farm located in a
Mediterranean Environment (Croatia). Environ Earth Sci 79:70
22. Cannell RQ (1977) Soil aeration and compaction in relation to root growth and soil manage-
ment. Appl Biol 2:1–86
23. Zhang Y, Li P, Liu XJ, Xiao L, Shi P, Zhao BH (2019) Effects of farmland conversion on the
stoichiometry of carbon, nitrogen, and phosphorus in soil aggregates on the Loess Plateau of
China. Geoderma 351:188–196
24. Charlotte V, Laure VG, Naoise N, Claire C (2022) Opportunities and limits in imaging
microorganisms and their activities in soil microhabitats. Soil Biol Biochem 174:108807
25. Gao CM, Wang M, Ding L, Chen YP, Lu ZF, Hu J, Guo SW (2020) High water uptake ability
was associated with root aerenchyma formation in rice: evidence from local ammonium supply
under osmotic stress conditions. Plant Physiol Biochem 150:171–179
26. Hao JQ, Lin Y, Ren GX, Yang GH, Han XH, Wang XJ, Ren CJ, Feng YZ (2021) Compre-
hensive benefit evaluation of conservation tillage based on BP neural network in the Loess
Plateau. Soil Tillage Res 205:104784
27. Praveena AAA, Tamilnesan P, Muthukumaran M, Udayakumar MD (2021) Experimental
analysis of moisture content with involuntary irrigation structure in soil. Mater Today 45:
1893–1897
28. Pi HW, Huggins DR, Sharratt B (2019) Dry aggregate stability of soils influenced by crop
rotation, soil amendment, and tillage in the Columbia Plateau. Aeolian Res 40:65–73
29. Ma PP, Nan SZ, Yang XG, Qin Y, Ma T, Li XL, Yu Y, Bodner G (2022) Macroaggregation is
promoted more effectively by organic than inorganic fertilizers in farmland ecosystems of
China—A meta-analysis. Soil Tillage Res 221:105394
30. Jabro JD, Stevens DB, Iversen WM, Man Sainju U, Allen BL (2021) Soil cone index and bulk
density of a sandy loam under no-till and conventional tillage in a corn-soybean rotation. Soil
Tillage Res 206:104842
31. Gura I, Mnkeni PNS (2019) Crop rotation and residue management effects under no till on the
soil quality of a Haplic Cambisol in Alice, Eastern Cape, South Africa. Geoderma 337:927–934
32. Lv LG, Gao ZB, Liao KH, Zhu Q, Zhu JJ (2023) Impact of conservation tillage on the
distribution of soil nutrients with depth. Soil Tillage Res 225:105527
33. Hua RW, Liu YJ, Chen T, Zheng ZY, Peng GJ, Zou YD, Tang CG, Shan XH, Zhou QM, Li J
(2021) Responses of soil aggregates, organic carbon, and crop yield to short-term intermittent
deep tillage in Southern China. J Clean Prod 298:126767
34. Wellingthon SGJ, Etienne D, Isabella CM, Cezar FAJ, Camila VVF, Zigomar MS (2019)
Prediction of soil stresses and compaction due to agricultural machines in sugarcane cultivation
systems with and without crop rotation. Sci Total Environ 681:424–434
35. Janulevičius A, Damanauskas V, Pupinis G (2018) Effect of variations in front wheels driving
lead on performance of a farm tractor with mechanical front-wheel-drive. J Terramech 77:23–30
Agricultural Land Degradation in China 173

36. Jin LQ, Peng XL, Liu J, Zhang QX, Li JH, Li LG (2022) Robust algorithm of indirect tyre
pressure monitoring system based on tyre torsional resonance frequency analysis. J Sound Vib
538:117198
37. Alex K, Nigel H, Peeyush S, Madhav DG, Simon C, Jannatul F (2013) A review of the tractive
performance of wheeled tractors and soil management in lowland intensive rice production. J
Terramech 50:45–62
38. Ten Damme L, Stettler M, Pinet F, Vervaet P, Keller T, Juhl ML, Lamandé M (2019) The
contribution of tyre evolution to the reduction of soil compaction risks. Soil Tillage Res
194:104283
39. Lucero M, Scott XC, Richard K (2006) Effects of tree harvesting, forest floor removal, and
compaction on soil microbial biomass, microbial respiration, and N availability in a boreal
aspen forest in British Columbia. Soil Biol Biochem 38:1734–1744
40. Donner J, Bravo S, Stoorvogel M, Dec D, Valle S, Clunes J, Horn R, Uteau D, Wendroth O,
Lagos L, Zúñiga F (2022) Short-term effects of compaction on soil mechanical properties and
pore functions of an Andisol. Soil Tillage Res 221:105396
41. Yang C, Sun J (2021) Impact of soil degradation on plant communities in an overgrazed Tibetan
alpine meadow. J Arid Environ:104586
42. Li SG, Harazono Y, Oikawa T, Zhao HL, He ZY, Chang XL (2000) Grassland desertification by
grazing and the resulting micrometeorological changes in Inner Mongolia. Agric For Meteorol
102:125–137
43. Du CJ, Gao YH (2021) Grazing exclusion alters ecological stoichiometry of plant and soil in
degraded alpine grassland. Agric Ecosyst Environ 308:107256
44. Saheed OJ, Feng X, Li P, Hou YL, Hou XY (2020) Risk-overgrazing relationship model: an
empirical analysis of grassland farms in Northern China. Rangel Ecol Manag 73:463–472
45. Zhou J, Zhong H, Hu WY, Qiao GH (2022) Substitution versus wealth: dual effects of
non-pastoral income on livestock herd size. World Dev 151:105749
46. Wang SJ, Zuo QQ, Cao QB, Wang P, Yang B, Zhao S, Cao R, Chen MK (2021) Acceleration of
soil N2O flux and nitrogen transformation during tropical secondary forest succession after
slash-and-burn agriculture. Soil Tillage Res 208:104868
47. Jakelyne SB, Víctor AR, Jonathan MT, Adrielle L, Inara RL, Marcelo T (2022) Drastic
impoverishment of the soil seed bank in a tropical dry forest exposed to slash-and-burn
agriculture. For Ecol Manag 513:120185
48. Wang SJ, Chen MK, Cao R, Cao QB, Zuo QQ, Wang P, Yang B, Zhao S (2020) Contribution of
plant litter and soil variables to organic carbon pools following tropical forest development after
slash-and-burn agriculture. Land Degrad Dev 30:1071–1077
49. Serrani D, Cocco S, Cardelli V, D’Ocatvio P, Rafael RBA, Domingos F, Vilanculos A,
Fernández-Marcos M, Giosué C, Tittarelli F, Corti G (2022) Soil fertility in slash and burn
agricultural systems in central Mozambique. J Environ Manag 322:116031
50. Selvaraj S, De la Rosa J, Huang ZJ, Guo FT, Ma XQ (2018) Effects of ageing and successive
slash-and-burn practice on the chemical composition of charcoal and yields of stable carbon.
Catena 162:141–147
51. Lu M, Wang SJ, Zhang Z, Chen MK, Li SH, Cao R, Cao QB, Zuo QQ, Wang P (2019)
Modifying effect of ant colonisation on soil heterogeneity along a chronosequence of tropical
forest restoration on slash-burn lands. Soil Tillage Res 194:104329
52. Hou XH, Xiang YZ, Fan JL, Zhang FC, Hu WH, Yan FL, Xiao C, Li YP, Cheng HL, Li ZJ
(2022) Spatial distribution and variability of soil salinity in film-mulched cotton fields under
various drip irrigation regimes in southern Xinjiang of China. Soil Tillage Res 223:105470
53. Zhang YT, Hou K, Qian H, Gao YY, Fang Y, Xiao S, Tang SQ, Zhang QY, Qu WG, Ren WH
(2022) Characterization of soil salinisation and its driving factors in a typical irrigation area of
Northwest China. Sci Total Environ 837:155808
54. Ke ZM, Liu XL, Ma LH, Feng Z, Tu W, Dong QG, Jiao F, Wang ZL (2021) Rainstorm events
increase risk of soil salinisation in a loess hilly region of China. Agric Water Manag 256:107081
174 Y. Yu et al.

55. Zhao YT, Wang GD, Zhao ML, Wang M, Jiang M (2022) Direct and indirect effects of soil
salinisation on soil seed banks in salinising wetlands in the Songnen Plain, China. Sci Total
Environ 819:152035
56. Zhuang QW, Shao ZF, Huang X, Zhang Y, Wu WF, Feng XX, Lv XW, Ding Q, Cai BW,
Altand O (2021) Evolution of soil salinisation under the background of landscape patterns in the
irrigated northern slopes of Tianshan Mountains, Xinjiang, China. Catena 206:105561
57. Ma LG, Ma FL, Li JD, Gu Q, Yang ST, Wu D, Feng J, Ding JL (2017) Characterizing and
modeling regional-scale variations in soil salinity in the arid oasis of Tarim Basin, China.
Geoderma 305:1–11
58. Wang QM, Huo ZL, Zhang LD, Wang JH, Zhao Y (2016) Impact of saline water irrigation on
water use efficiency and soil salt accumulation for spring maize in arid regions of China. Agric
Water Manag 163:125–138
59. Cui H, Bai JH, Du SD, Wang JJ, Ghemelee NK, Wang W, Zhang GL, Jia J (2021) Interactive
effects of groundwater level and salinity on soil respiration in coastal wetlands of a Chinese
delta. Environ Pollut 286:117400
60. Yu L, Tao Z, Bai JH, Wang JJ, Yu ZB, Wang X, Zhang GL (2020) Effects of water and salinity
on soil labile organic carbon in estuarine wetlands of the Yellow River Delta, China. Ecohydrol
Hydrobiol 20:556–569
61. Cheng ZB, Wang JY, Gale WJ, Yang HC, Zhang FH (2020) Soil aggregation and aggregate-
associated organic carbon under typical natural halophyte communities in arid saline areas of
Northwest China. Pedosphere 30:236–243
62. Salcedo FP, Cutillas PP, Alarcón Cabañero JJ, Gaetano Vivaldi A (2022) Use of remote sensing
to evaluate the effects of environmental factors on soil salinity in a semi-arid area. Sci Total
Environ 815:152524
63. Carvalho FP (2017) Pesticides, environment, and food safety. Food Energy Secur 6:48
64. Guo AY, Bao YY, Zhou QX (2020) Advances in soil pesticide contamination and bacterial
pesticide-antibiotic cross-resistance. Microbiol China 47:2984–2995
65. Xu P, Feng YP, Fan J, Liu HP, Lu BW, Liu F, Chen HH (2014) Organochlorine pesticides
pollution in soils of typical areas in China: recent advances and future prospects. Agrochemicals
53:164–166
66. Lin BL, Meng YB, Kamo M, Naito W (2021) An all-in-one tool for multipurpose ecological
risk assessment and management (MeRAM) of chemical substances in aquatic environment.
Chemosphere 268:128826
67. Yang YH, Chen TT, Liu XC, Wang S, Wang K, Xiao R, Chen XP, Zhang T (2022) Ecological
risk assessment and environment carrying capacity of soil pesticide residues in vegetable
ecosystem in the Three Gorges Reservoir Area. J Hazard Mater 435:128987
68. Badrzadeh N, Vali Samani JM, Mazaheri M, Kuriqi A (2022) Evaluation of management
practices on agricultural non-point source pollution discharges into the rivers under climate
change effects. Sci Total Environ 838:156643
69. Chen YF, Liu DX, Ma JH, Jin BY, Peng JB, He XL (2021) Assessing the influence of
immobilisation remediation of heavy metal contaminated farmland on the physical properties
of soil. Sci Total Environ 781:146773
70. Lin H, Wang ZW, Liu CJ, Dong YB (2022) Technologies for removing heavy metal from
contaminated soils on farmland: a review. Chemosphere 305:135457
71. Zhou B, Li X (2021) The monitoring of chemical pesticides pollution on ecological environ-
ment by GIS. Environ Technol Innov 23:101506
72. Naidu R, Biswas B, Willett IR, Cribb J, Kumar Singh KS, Nathanail CP, Coulon F, Semple KT,
Jones KC, Barclay A, Aitken RJ (2021) Chemical pollution: a growing peril and potential
catastrophic risk to humanity. Environ Int 156:106616
73. Aragón MA, Mitsui Nakamaru Y, García-Carmona M, Martínez Garzón FG, Martín Peinado FJ
(2019) The role of organic amendment in soils affected by residual pollution of potentially
harmful elements. Chemosphere 237:124549
Agricultural Land Degradation in China 175

74. Kumar M, Xiong XN, He MJ, Tsang DCW, Gupta JH, Khan E, Harrad S, Hou DY, Ok YS,
Bolan NS (2020) Microplastics as pollutants in agricultural soils. Environ Pollut 265
(Pt A):114980
75. Xu XB, Sun MX, Zhang LX, Xue YH, Li C, Ma SY (2021) Research progress and prospect of
soil microplastic pollution. J Agric Resour Environ 38:1–9
76. Nizzetto L, Futter M, Langaas S (2016) Are agricultural soils dumps for microplastics of urban
origin? Environ Sci Technol 50:10777–10779
77. Chai BW, Wei Q, She YZ, Lu GN, Dang Z, Yin H (2020) Soil microplastic pollution in an
e-waste dismantling zone of China. Waste Manag 118:291–301
78. Li J, Song Y, Cai YB (2020) Focus topics on microplastics in soil: analytical methods,
occurrence, transport, and ecological risks. Environ Pollut 257:113570
79. Mahon AM, O’Connell B, Healy MG, O’Connor I, Officer R, Nash R, Morrison L (2017)
Microplastics in sewage sludge: effects of treatment. Environ Sci Technol 51:810–818
80. Li XW, Chen LB, Me QQ, Dong B, Dai XH, Ding GJ, Zeng EY (2018) Microplastics in sewage
sludge from the wastewater treatment plants in China. Water Res 142:75–85
81. Klein M, Fischer EK (2019) Microplastic abundance in atmospheric deposition within the
Metropolitan area of Hamburg, Germany. Sci Total Environ 685:96–103
82. Guo JJ, Huang XP, Xiang L, Wang YZ, Li YW, Li H, Cai QY, Mo CH, Wong MH (2020)
Source, migration and toxicology of microplastics in soil. Environ Int 137:105263
83. Gao B, Yao HY, Li YY, Zhu YZ (2021) Microplastic addition alters the microbial community
structure and stimulates soil carbon dioxide emissions in vegetable-growing soil. Environ
Toxicol Chem 40:352–365
84. Chae Y, An YJ (2018) Current research trends on plastic pollution and ecological impacts on
the soil ecosystem: a review. Environ Pollut 240:387–395
85. Wright RJ, Bosch R, Langille MGI, Gibson MI, Christie-Oleza JA (2021) A multi-OMIC
characterisation of biodegradation and microbial community succession within the PET
plastisphere. Microbiome 9:141
86. Mintenig SM, Int-Veen I, Löder MGJ, Primpke S, Gerdts G (2017) Identification of
microplastic in effluents of waste water treatment plants using focal plane array-based micro-
Fourier-transform infrared imaging. Water Res 108:365–372
87. Ren XW, Tang JC, Yu C, He J (2018) Advances in research on the ecological effects of
microplastic pollution on soil ecosystems. J Agro-Environ Sci 37:1045–1058
88. Talvitie J, Mikola A, Koistinen A, Setälä O (2017) Solutions to microplastic pollution –
removal of microplastics from wastewater effluent with advanced wastewater treatment tech-
nologies. Water Res 123:401–407
89. Sun J, Dai XH, Wang QL, van Loosdrecht MCM, Ni BJ (2019) Microplastics in wastewater
treatment plants: detection, occurrence and removal. Water Res 152:21–37
Agricultural Soil Degradation in Colombia

Mauricio Quintero-Angel and Daniel I. Ospina-Salazar

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
2 Natural Regions of Colombia and Taxonomic Diversity of Soils . . . . . . . . . . . . . . . . . . . . . . . . . 179
3 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
4.1 Agricultural Soil Degradation Process in Colombia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
4.2 Soil Degradation Extension in Natural Regions of Colombia . . . . . . . . . . . . . . . . . . . . . . . 197
4.3 Main Causes and Consequences of Soil Degradation in Colombia . . . . . . . . . . . . . . . . . 199
4.4 Conflicts by Soil Use in Colombia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
4.5 Viable Agricultural Practices to Reduce Soil and Land Degradation . . . . . . . . . . . . . . . 205
4.6 Research Gaps and Enabling Factors in Colombia for Soil and Land Sustainability 207
5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213

Abstract For Colombia, reversing land and soil degradation is a fundamental


challenge looking to guarantee food security and conservation of its wide ecosystem
diversity. This chapter analyzes trends and challenges of land and soil degradation
associated with agricultural practices in Colombia. Using a methodological frame-
work based on expert consultation, systematic search of information, and
bibliometric analysis, we ask about the importance of different degradation process
on agricultural land, as well as their main causes and its problematics. At a regional
level we analyze the extension of different processes of land and soil degradation and
agricultural practices which can or will be more viable to reduce it. Furthermore, we

M. Quintero-Angel (✉)
Universidad del Valle-Sede Palmira, Palmira, Colombia
e-mail: [email protected]
D. I. Ospina-Salazar
Universidad del Valle-Sede Palmira, Palmira, Colombia
Departamento de Ciencias Agronómicas, Universidad Nacional de Colombia, Medellín,
Colombia
e-mail: [email protected]

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 177
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 177–218, DOI 10.1007/698_2022_914,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 22 December 2022
178 M. Quintero-Angel and D. I. Ospina-Salazar

identify research gaps or current problems which demands immediate and effective
actions to advance towards a sustainable management of soils and lands. Finally, we
conclude that to stop agricultural and land degradation in Colombia, and, if possible,
to reverse this process, is only feasible through state-level policies and the transfor-
mation of paradigms in which diverse actors impact soils and land with their
agricultural activities.

Keywords Agricultural frontier, Cattle ranching, Research gaps, Soil erosion,


Sustainability

1 Introduction

Due to its location in the northwestern corner in South America, Colombia has two
of the most important physiographic units of the subcontinent: Andes range and
Amazon basin. Both sum to volcanic activity and weathering of diverse parent
materials has been resulted on a high diversity of soils, which is threatened since
the twentieth century.
Degradation processes such as water erosion, loss of organic matter, soil com-
paction, micro and macro fauna reduction and agrochemical pollution, among others
had affected gravely the soils of the country [1]. However, intensity and extension of
degradation processes are not evenly distributed, it varies depending on the region1
and socio-economic activities in the territory; e.g., water and wind erosion, soil
compaction, and loss of organic matter predominantly occur in the Caribbean and
Andes regions. Currently, more than 40% of continental lands of the country are
affected by erosion, 5% by salinization and 24% are susceptible to desertification
[3]. Other forms of soil degradation, such as pollution with agrochemical, heavy
metals, and micro plastics, are less studied.
There are multiple factors that influence soil and land degradation in Colombia.
Urban sprawl is responsible for permanent soil loss, while socio-economic activities,
namely agro-industrial modernization, had generated multiple changes in the soil
uses, leading to transformation of rain forests, tropical dry forest, natural savannas,
swamps, and moorlands in croplands, pastures, and/or timber plantations. Moreover,
Colombian economic model in the last years is oriented to reprimarisation of the
economy, which had led to an increase of mining extractive projects with irreversible
negative impacts in the soil.
In Colombia there is a limited knowledge and assessment of soil, its function and
conservation importance, which is product of accumulated weakness in national
education, science, technology, and innovation systems [1]. Moreover, in the last
years, the Colombia had experience a complex socio-political situation in which

1
Colombia is divided into six natural regions: Caribbean, Pacific, Andean, Orinoquian, Amazonian,
and Insular [2]. In Colombia there are 32 departments, which constitute the administrative divisions
of the country.
Agricultural Soil Degradation in Colombia 179

confluence of high inequality, poverty, presence of illegal army groups, illicit


activities linked to narcotic production, and illegal exploitation of minerals which
negatively affect adequate management of soils [1]. In this challenging context, this
chapter aims to analyze trends and challenges of land and soil degradation associated
to agricultural practices in Colombia, as well as its causes and more relevant
consequences.

2 Natural Regions of Colombia and Taxonomic Diversity


of Soils

Colombia is located at the northwest of South America between 12° 30′ 46″ N and
4° 13′ 30″ S and 66° 50′ 54″ W and 79° 01′ 23″ W. The country has borders with
Venezuela and Brazil in the east, Ecuador and Peru in the south, the Caribbean Sea in
the north, northeast with Panama and the Pacific Ocean to the west [2]. It has a total
area of 2,070,408 km2, from which 1,141,748 km2 is continental area and
928,660 km2 is maritime surface [4]. The country is inhabited by 48,258,494 people,
mainly settled in cities, (77.1%), rural villages (7.1%), and dispersed rural areas
(15.8%) [5].
Diversity in topography, weather, vegetation, soils, and distance to sea results in
six natural regions [6] (Fig. 1). Andean region is composed of all territories on the
three mountain ranges of the Andes. This area concentrates most of the population
and socio-economic development. Caribbean and Pacific regions are located next to
the Caribbean Sea in the north and the Pacific Ocean to the west, respectively.
Amazon region owes its name to the Amazon Forest, in the southernmost part of the
country, while Orinoquian region encompasses the extended plains along the Ori-
noco River in the east. Amazonian and Orinoquian regions represent approximately
60% of the national territory and 5% of the population. Finally, insular region
includes San Andres and Providencia islands in the Caribbean and other islets and
keys in the Pacific [7].
Colombian climate is determined by its position at the tropical zone, the influence
of the intertropical convergence zone (ITCZ), as well as other geographic and
atmospheric features (precipitation, solar radiation, temperature, wind systems,
altitude, continentality, and atmospheric humidity). All these factors produce diverse
climates and microclimates conditions, from hot in the coast and plains up to cold in
the Andean mountains and Sierra Nevada of Santa Marta in the Caribbean
[4]. Annual precipitation in Colombia is 3,000 mm, with a real evapotranspiration
of 1,180 mm and mean annual runoff of 1,830 mm [4]. The lowest annual precip-
itation is observed in the northernmost extreme of Guajira peninsula region with
500 mm or less, while the highest is registered in the Pacific region with values of
more than 10,000 mm [4].
Colombian pedosphere contains 11 of the 12 taxonomic orders of soils according
to the USDA classification system, with exception of Gelisols. They mostly belong
to incipient and less-evolved soils, such as Entisols and Inceptisols, and acid, highly
weathered soils such as Ultisols and Oxisols (Table 1) [8].
180 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 1 Natural regions of Colombia

According to Instituto Geográfico Agustin Codazzi (Geographic Institute of


Colombia) [9], most of the soils suitable for agriculture matches with the distribution
of Mollisols, Andisols, and Alfisols, which covers 8.33% of the country. However,
coffee plantations are also raised on Entisols and Inceptisols [9]. Pastures for cattle
ranching are distributed on almost all of the soil orders but dominantly cover Oxisols
and Ultisols at the Orinoquian region.
Agricultural Soil Degradation in Colombia 181

Table 1 Soil orders present in Colombia


Extension in hectares Departments and
(% of national formations where can be
Soil order Characteristics territory) found
Alfisols Base saturation percentage 985,655 (0.86%) Dry and flat areas close to
of at least 35%, presence of river valleys
horizons with clay accumu-
lation forming hard layers,
developed under dry to
sub-humid, temperate cli-
mates. High natural fertility
for agriculture
Andisols Soils derived from ash and 6,767,851 (5.93%) Mountain landscapes, more
volcanic debris, mineral extended in high plateaus,
fraction with predominance piedmont, and alluvial
of allophanes, imogolites, plains. Found in the
and ferrihydrites. Low bulk Andean region with pre-
density, great porosity, high dominance in the central
water holding capacity, high range
cation exchange capacity
(CEC) but phosphorus-
deficient
Aridisols Soils of arid and semi-arid 608,941 (0.53%) Specially located on
climates with presence of Guajira department and in
calcium carbonate, high very dry areas of Magda-
content of soluble salts, low lena River valley
content of organic matter
and presence of clay hori-
zons, gypsum, or salts
depositions by illuviation
during high precipitation
events
Entisols Soils with practically no 22,588,680 (19.78%) Distributed in all Colom-
development, without bian territory
apparent horizons for diag-
nostic. Typical in steep hill-
sides with deposition of
sediments, parent material
almost exposed, without
weathering. Present in
almost any environmental
conditions
Histosols Soils with thick organic 358,835 (0.31%) Located in river plains or
layer, little transformed, around swamps, marshes,
very low bulk density and and bogs in high areas of
high water holding capacity. the Andes, particularly in
Evolved from dried water Boyacá, Cundinamarca,
bodies, which expose the and Nariño departments
soil to aeration and alteration
processes. Soils highly sus-
ceptible to acidic and sulfur
accumulation processes,
which are difficult to reverse
(continued)
182 M. Quintero-Angel and D. I. Ospina-Salazar

Table 1 (continued)
Extension in hectares Departments and
(% of national formations where can be
Soil order Characteristics territory) found
Inceptisols Incipient soils, although not 43,758,312 (38.33%) Distributed in all Colombia,
as much as Entisols, are in association with other
characterized by low devel- soil orders
opment of diagnostic hori-
zons. Frequently have more
evident coloration or clay
content, but without accu-
mulation layers due to illu-
viation. Occurs almost in
all-climates but dominantly
in arid ones. Any soil that is
not possible to be classified
in any of the other orders is
considered an Inceptisol
Mollisols Soils from semi-arid and 1,752,929 (1.54%) Located on Cauca, Sinú,
semi-humid environments, San Jorge, and Magdalena
deep, with presence of River valleys
mollic horizon, and high
content of transformed
organic matter (humus).
Very suitable for agriculture,
they show base saturation of
at least 50%
Oxisols Highly weathered, 14,941,577 (13.09%) Located mainly in
red-yellowish soils with Orinoquian and Amazonian
high content of free soluble regions
aluminum (Al3+), acidic,
poor fertility and CEC.
Mainly present in humid and
hot tropical climates with
mild slopes, having good
drainage due to its high
porosity
Spodosols Naturally infertile soils, 921,162 (0.81%) Located mainly in Guainía
podzolized and developed in and Vaupés departments
very humid environments. It (Amazonian region)
presents a horizon of organic
matter associated to alumi-
num and iron by illuviation
(spodic horizons), under a
sandy, deep horizon of ligh-
ter color (ash gray). Low pH
and clayey textures, not
suitable for agriculture
Ultisols Highly weathered soils, 17,921,151 (15.70%) Located mainly on
acidic, with low fertility, Orinoquian, Amazonian,
saturated most of the year, and Pacific Coast regions
with high content of iron
oxides, showing a clay
accumulation horizon
(continued)
Agricultural Soil Degradation in Colombia 183

Table 1 (continued)
Extension in hectares Departments and
(% of national formations where can be
Soil order Characteristics territory) found
Vertisols Soils with high content of 851,507 (0.75%) Located in the Caribbean
montmorillonites clays, region and especially in the
prone to cracking, high con- Valle del Cauca department
tent of organic matter, fertile
for many crops, short time
for optimal tillage. Charac-
teristic of humid environ-
ments with contrastable dry
and humid seasons
Based on information from Instituto Geográfico Agustín Codazzi [8]

3 Methodology

To identify the main problems of land and soil degradation in different natural
regions of Colombia, a consultation with experts on agricultural soils and similar
areas was conducted. The experts were selected from a list of national experts of soil
science and a snowball sampling or chain-referral sampling conducted during
January and February 2021. The expert consultation was anonymized and followed
the American Anthropological Association Statement on Ethics. Consultation with
experts is a scientifically validated method that employs key informants with
experience in the topics of interest to quickly obtain knowledge related to a particular
topic [10]. Experts were asked to voluntarily answer several questions through
Google Forms. The form inquired about some characteristics of experts, including
gender, higher education, and years of experience. It also contains questions about
importance of different agricultural land and soil degradation processes in Colombia.
Additionally, it inquired the main causes and derived consequences of land degra-
dation and Science, Technology, and Innovation gaps in Colombia related to
agricultural soil and land degradation. For each natural region, where the expert’s
professional experience was concentrated, experts were asked for the extension of
land and soil degradation, the agricultural practices which produce the main degra-
dation, and practices for reversing it. From quantitative information, reported by
experts a descriptive statistic was made. Moreover, qualitative information was
organized deductively by categories to get patterns. Answers from each expert
were organized by topic codifying the source of anonymous information and were
ordered in an analysis matrix.
Additionally, a systematic search of information was carried out Spanish and
English on soil and land degradation, according to some search criteria, information
sources, and predefined keywords, along with the three basic Boolean operators
AND, OR, and NOT, as described below: (1) Search criteria: Policies, reports,
articles, and scientific publications on soil or land degradation in Colombia;
184 M. Quintero-Angel and D. I. Ospina-Salazar

(2) Information sources: meta search engines Google, Google Scholar, and Scopus;
(3) Keywords: soil, land, degradation, Colombia, erosion, organic matter loss,
compaction, desertification, micro and macro fauna reduction, salinization, sodicity,
acidification, soil fertility loss, and contamination. Specific search in Scopus was
made with the following algorithm: TITLE-ABS-KEY (land OR soil AND degra-
dation AND Colombia) AND (LIMIT-TO (DOCTYPE, “ar”) OR LIMIT-TO
(DOCTYPE, “re”) OR LIMIT-TO (DOCTYPE, “ch”). To track the thematic evolu-
tion of scientific publications on land and soil degradation in Colombia, the results of
Scopus were processed in SciMAT (Science Mapping Analysis Software Tool)
[11]. The latter in order to cluster keywords and their interconnections in research
themes to generate a strategic diagram, which is a bidimensional space to graph
themes according to their centrality and density values [12].

4 Results and Discussion


4.1 Agricultural Soil Degradation Process in Colombia

Expert consultation was voluntarily answered by 81 experts (59 males and


22 females). The 40.7% (33) of experts reported a doctoral degree, 33.3% (27) a
master’s degree, 9.9% (8) a specialization degree, and 16% (13) a bachelor’s degree,
as their higher education level achieved. Experts reported formation in disciplines
such as agronomy, agrology, biology, microbiology, chemistry, geology, geogra-
phy, zootechnics, forest engineering, agricultural engineering, agroforestry, and
agroecology. In terms of their years of experience related to soils, 69.1% (56) of
experts reported more than 10 years, 14.8% (12) reported between 5 and 10 years,
11.1% (9) between 3 and 5%, and 4.9% (4) less than 3 years.
In relation to the agricultural soil degradation in Colombia, experts agree that the
main degradation processes are organic matter loss (81 experts), water and wind
erosion (80 experts), and soil compaction (79 experts). Desertification, micro and
macrofauna reduction, salinization and sodicity, acidification, loss of soil fertility,
and contamination were classified mainly as important and very important. For some
experts, the latter processes are less important at the national level or are not a
problem at all (Fig. 2).

4.1.1 Erosion

The most important challenge for soil conservation in Colombia is degradation by


erosion, resulting mainly from runoff, and at less extent by wind, taking place in
croplands and pastures with unsuitable agricultural management across the country.
Erosion is a natural process carried on by water, wind and gravity. However,
degradation by erosion occurs when anthropic actions accelerate and/ or magnify
Agricultural Soil Degradation in Colombia 185

Fig. 2 Expert perception on the importance of soil degradation processes in Colombia

erosion above soil formation rate. This generates a net loss of soil, affecting its
functions and ecosystem services and reducing its productive capacity [13].
Unlike other more specific degradation processes, such as salinization and desert-
ification, erosion occurs along all natural’s regions in magnitudes and severities that
affect large portions of the territory. Forty percent of the soils of the continental and
insular area of the country (45,379,057 ha) are affected by some degree of erosion as
follows; 2.9% (3,334,594 ha) show severe and very severe erosion, 16.8%
(19,222,575 ha) moderate erosion and 20% (22,821,889 ha) mild erosion [13].
Water erosion is the predominant type, showed mainly as of laminar class
(19.33%), furrows (9.31%), terraces (7.30%), and gullies (0.52%), in practically
the entire country. Wind erosion is manifested in less than 1% of the territory,
particularly in the Guajira desert, as deflation depressions, desert pavements, and
dunes [13]. In Colombia, there are 34 hotspots2 of erosion, concentrated in the
Andean and Caribbean regions (Fig. 3).
The severity degree of erosion varies according to the soil cover, being higher in
cropland and pastureland, and very low in woodlands. Likewise, croplands have
greater susceptibility to erosion compared to pastureland, with the highest erosion
values occurring in cultivated hillsides. Although the extent of moderate erosion is
greater in croplands, the extent of severe degree is lower, in comparison with soils
under pastures (Fig. 4). Several authors report that, in this soil cover, the net loss of
soil is usually greater than in pastures, due to the frequent management of crops
without weed cover [14–17]. However, in pasture soils runoff is higher than in
cultivated soils due to a lower infiltration capacity, which can produce severe
overland flow under heavy rainfall [15, 18, 19].

2
A place of significant activity or danger.
186 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 3 Hotspots of erosion in Colombia. Based on information from Instituto de Hidrología,


Meteorología y Estudios Ambientales and Universidad de Ciencias Aplicadas y Ambientales [13]

Soil loss (kg ha-1 year-1) in Colombia has been estimated experimentally or by
modeling, taking into account the soil cover and the rainfall intensity (Table 2). As a
reference value in humid tropical regions, it is estimated an annual loss of
15,570 kg ha-1 in bare soil, 825 kg ha-1 in soils covered by crops, 374 kg ha-1 in
grasslands, while in forest soils it reaches 107 kg ha-1 [16].
Agricultural Soil Degradation in Colombia 187

Fig. 4 Severity degree of erosion by soil cover in Colombia. Elaboration from information of
Instituto de Hidrología, Meteorología y Estudios Ambientales andAmbientales and Universidad de
Ciencias Aplicadas y Ambientales [13]

Table 2 Erosion soil values for different land covers in Colombia


Crop without Natural Time
cover Pastures vegetation lapse Source
Soil loss values 1,839 1,091 1,170 Annual [17]
(kg ha-1) 40,000–75,000 2,200–1,500a – Annual [20]
2,600 200 200 Biannual [14]
2,300 200 – Biannual [21]
70–620 – – 18 months [22]
700–14,400b 700–14,400b 700–2,700 Annual [23]
10,400 500 – Annual [24]
a
Vetiver grass
b
Slope-dependent

The degradation by erosion associated with agricultural activities not only


reduces the availability of nutrients in the soil, but also creates the conditions for
mass movements (very frequent in the rainy season in Colombia) and increases the
sediment load in rivers, causing a general disturbance in socio-ecosystems. In
addition, it generates a differential socio-economic impact according to the position,
with the higher socio-economic effects on small owners, familiar production units,
peasants, and ethnic communities [25].
188 M. Quintero-Angel and D. I. Ospina-Salazar

4.1.2 Compaction

Degradation by compaction has been studied using indicators such as bulk density,
porosity, infiltration rate, and resistance to penetration, mostly in pasturelands, and
to a lesser extent in croplands. The available literature confirms that cattle grazing is
the main cause of soil degradation by compaction in Colombia.
Induced compaction by grazing is well described from the agricultural frontier in
Guaviare department [26]. Here, slash-and-burn agriculture is practiced in the
predominant Ultisols, taking advantage of the nutrient-rich ash, to initially fertilize
corn and cassava. Subsequently, grasses (Brachiaria decumbens) are sown with very
high planting distances, which leaves large portions of bare soil while the grass
covers the entire area. The livestock cycle begins with very low densities (less than
one cow per hectare), which can be managed for up to a decade before the pastures
degrade and get outplaced by weeds. As soil quality indicators, bulk density
increased 28%, as well as the resistance to penetration (ten times higher), while
total porosity and infiltration rate were reduced drastically, with respect to an intact
soil of the surrounding forest, in an evaluation from 3 to 9 years. The increase in bulk
density was negatively related to the biomass and crude protein yield of the
established pastures. However, as long as land is not scarce in these regions, soils
degraded by these changes in land use do not arouse interest in being recovered for
further agricultural exploitation [26].
The effect of livestock trampling is considerable in the compaction of soils with
isohyperthermic and high humidity regime, typical of Orinoquian and Amazon
regions. In hilly and low terrace soils with Brachiaria decumbens coverage for cattle
ranching in the last 15 years, notable increases in resistance to penetration are
reported in the first 15 cm of depth, alongside to a reduction of up to ten times in
infiltration rate [27]. Similarly, in two soils under pastures of the foothills of Caquetá
department, a reduction in total porosity, an increase in bulk density, and in
resistance to penetration (<3.0 MPa) is reported in the first 30 cm, compared to a
soil with intact native vegetation [28].
In the Andean region, agricultural soils showed significant compaction and
modification of the soil-air-water relationship, 2 years after the establishment of
the pastures [29]. Soils historically devoted to livestock in the Caribbean region,
established in highly degraded tropical dry forest ecosystems, show erosion signs on
80–100% of the pastures, along with increased compaction and low nutrient avail-
ability. This is ascribed to overgrazing and prevalence of grasses, not taking into
account the diversity of native trees and bushes that could contribute to nutrient
cycling [30]. As a consequence of this, these soils exhibit high levels of compaction
indicated by penetration resistance (4.42–7.49 MPa) and bulk density
(1.33–1.36 g cm-3) [31].
Soils subjected to mechanized tillage have also undergone compaction processes,
namely, in sugar cane monocultures in Valle del Cauca (inter-Andean zone), where
agricultural machinery is frequently used on wet soils [32]. A laboratory compaction
experiment in two soils (Haplustol and Calciustol) at two moisture regimes showed
Agricultural Soil Degradation in Colombia 189

that, in these crops, the bulk density, porosity, and hydraulic conductivity were
drastically affected compared to an intact forest soil. Likewise, the higher content of
organic matter per se did not prevent the increase in compaction in very humid
soil [32].
Given that these soils are susceptible to compaction by agricultural machinery,
the Sugar Cane Research Center-CENICAÑA has established the most suitable
types of tires, wagons, axles and harvest lift to mitigate this degradation process
[33–35]. This could explain why some soils under mechanized sugarcane production
exhibit low values of bulk density (1.35 g cm-3) and a high infiltration rate, leading
to an optimal quality index for intensive agriculture, which suggests good practices
over time [36].

4.1.3 Reduction of Soil Organic Matter (SOM)

The reduction of SOM is initially the result of tillage incorporation of the above-
ground biomass, from which the organic matter is oxidized in the form of CO2. Over
time, the SOM content tends to stabilize or decrease as a function of the incorpora-
tion and mineralization of crop residues, as well as the respiration of the microbial
biomass. The dynamics of SOM in livestock systems is more complex, since it
depends on the type and intensity of rotation and the animal contribution of manure/
urine. As a degradation process, SOM reduction is closely linked to soil erosion and
deforestation, therefore, it manifests with greater intensity in tilled soils and pastures,
especially located in agricultural frontier regions.
In Andean areas above 2,000 m.a.s.l., dairy farming is often extended on mild to
harsh slopes. In these areas, low conservation of SOM is reported under intensive
management, due to the removal of the vegetation cover, plowing, and the low
contribution of organic matter, which increases the loss of carbon due to secondary
mineralization and surface runoff. However, in wild pastures, higher CO2 emissions
were noted due to the greater contribution of organic matter subjected to minerali-
zation by microorganisms [37].
The complexity of soil organic carbon (SOC) and its interchangeability between
different pools (residual carbon, microbial biomass, humus) makes it difficult to
predict management practices that steady increases SOC in degraded soils. For
example, the establishment of trees like Albizia saman and Guazuma ulmifolia did
not significantly increase SOC in silvopastoral systems, when compared to that of a
degraded pasture [38]. This may be due to the dynamics of the litter that falls from
the trees, which determines the tradeoff of organic matter into the soil and the
intensity of the nutrient cycle [38]. In Amazonian pastures with low grazing inten-
sity, SOC tends to increase 20 years after the deforestation of the land, even above
the level before deforestation. On the contrary, in pastures with intensive grazing,
SOC suffered a decrease of up to 27% compared to the soil prior to deforestation
[39]. The latter results contrast with the increase in SOC in an intensive, highly
rotational grazing system, which was attributed to a higher density of grass
roots [40].
190 M. Quintero-Angel and D. I. Ospina-Salazar

4.1.4 Acidification

Colombian soils are generally acid due to dissolving of acid parent material and
humid climate conditions that characterize great part of its territory. It is estimated
that 85.6% of the soils have a pH lower than 5.5 [41]. However, most of these soils
are used in agricultural systems. Without liming, these soils tend to undergo acid-
ification, especially in the Orinoquian region [42]. The foregoing does not imply that
this problem is inexistent in other regions; e.g., in 187,000 soil samples analyzed
from 11 Andean coffee-producer departments, 56% had pH values below 5.0 and
45% had Al3+ saturation percentages above 1% [43].
Acidification of soils can also be generated by the addition of acid reaction
fertilizers such as urea, diammonium phosphate, phosphoric acid or chlorides, as
well as by ammonium nitrification and lixiviation of exchangeable bases and their
replacement by aluminum. The exchange of bases/H+ in the root-soil system, the
decomposition of organic matter, and the respiration of the roots also generate soil
acidification. In Risaralda department (Andean region), lower pH values have been
reported in native forest soils than under pastures, unshaded-coffee and pine plan-
tations, due to the decomposition of organic matter that releases carboxylic acids and
phenolic compounds. In contrast, onion crop soils showed a slight reduction in pH
after 4 months of evaluation, together with a greater activity of ammonium and
nitrite-oxidizing bacteria [44]. The latter is interesting since pine plantations in
Colombia have been blamed for soil acidification and is a common believe among
people.
Other acidification processes have been detected as a result of the deposition of
volcanic materials after the eruption of Ruiz volcano in 1986. This type of acidifi-
cation is more pronounced in high altitude alpine-like soils than in plantation soils of
the surroundings [45]. Acid sulfate soils are also of local concern in some areas of
the northeast Andes (Chicamocha irrigation district, Tundama valley) and the
Caribbean plains (lower Sinú valley) [46, 47]. Banana cultivation, which is of
great economic importance in the Urabá gulf coast (Caribbean region), is known
for promoting soil acidification. This occurs because of the release of H+ from the
roots to the soil, and also for the application of ammonium fertilizers. Within 2 years
of intensive banana cultivation in this region, an increase of Al3+ by 5% was noted
[48]. Although sugar cane cultivation has been acknowledged as a promoter of soil
acidification worldwide [49], there is no evidence of acidification by this crop in
Colombia, at least in Valle del Cauca region where it is most cultivated, aside from
few reports [50].

4.1.5 Desertification

When defining desertification as the advance of desert sands over fertile lands, it can
be said that in Colombia this phenomenon is marginal, and possibly only occurs in
very specific sectors of the Guajira desert (Caribbean region). However, there are
Agricultural Soil Degradation in Colombia 191

large portions of the national territory affected by desertification, if defined as the


degradation of the lands of arid, semi-arid and dry sub-humid zones. This is notable
as the reduction or loss of biological or economic productivity and complexity of
agricultural land, resulting from various factors, such as climatic variations and
human activities [51]. In Colombia, desertification has been determined by three
indicators: climate (precipitation/evapotranspiration ratio <0.65), vegetation cover
(presence of xerophytes) and soil moisture regime (udic, ustic or xeric) [52].
The dry area covers 245,342 km2 of the country, approximately 21.5% of
the national territory. The Orinoquian and Caribbean regions are the ones with the
largest extension of dry area with 94,096 and 91,522 km2, respectively. In the
Andean region, most of the dry enclaves have been formed by rain shadow effect,
where the mountain ranges prevent the passage of humid air, creating dry valleys and
semi-arid canyons. In the Orinoquian plains, the extension of dry areas is wide due to
the intensity and duration of the dry season, which causes a very low precipitation/
evapotranspiration ratio. Almost 79% (193,510 km2) of the dry areas show some
level of degradation derived from erosion and salinization processes [53].
An illustrative case of the advance of desertification is that of the Cesar River
valley in the Caribbean. Between the 50s and 80s, landowners sponsored by
government soft loans cleared large portions of the tropical dry forest to grow cotton
with intensive tillage and the incorporation of pre-emergence herbicides such as
Trifluralin [54]. Over few years, these practices promoted soil structural degradation,
evinced as emerging of infertile horizons, reduction of the water holding capacity
and deterioration of the soil ecosystem services, effects that are still evident
nowadays [54].
In other dry regions (Patía river valley and Sierra Nevada de Santa Marta),
overgrazing, combined with burning in the dry season and the introduction of exotic
grasses, has favored desertification processes. One of the most studied dry ecosys-
tems is the Tatacoa Desert in the Upper Magdalena valley, which undergoes severe
erosion processes, yet the tendency to desertification has been recently questioned.
The formation of this desert millennia ago was possible thanks to the climate regime
and soils, but recent anthropic factors, such as goat and cattle grazing, have accel-
erated this process, by restraining the development of plant cover [55]. However,
between 1987 and 2010, it has been observed an increase of vegetation acreage,
although the replacement dynamic between vegetation and degraded areas is highly
stressing for plants, resulting in a patchy distribution [55].
Historically, the effects of desertification have been deeper in the Guajira penin-
sula. This sub-region, subjected to erratic and scarce rainfall, is prone to droughts,
impacted by coal mining and has high levels of poverty, converging factors that
magnify soil degradation, to an extent that has caused famines within the Wayuu
native communities [56]. These communities inhabit the driest sub-regions of the
peninsula, an issue that strongly influences the infant mortality of those families that
do not own goods such as animals (mainly goats) and savings from normal times that
can be used for consumption or to buy food during drought [56].
192 M. Quintero-Angel and D. I. Ospina-Salazar

4.1.6 Salinization and Sodicity

In the continental and insular area of Colombia, 12.3% (14,041,883 ha) of the soils
present some degree of degradation due to mild to moderate and severe salinization
[57]. Salinization usually occurs either in natural depressions where surface evapo-
ration is greater than precipitation, or where the water table is very high and/or where
the parent material produces salts when it weathers. The limitations to productivity
of saline or saline-sodic soils are related to dispersion processes and lack of avail-
ability of Mg2+ and Ca2+, which reduces the stability of the aggregates and generates
a poor physical condition of the soil. High sodium concentration also affects plant
development due to its toxicity on most species [58].
Saline and saline-sodic soils are mainly located in the Caribbean and Andean
regions, along Cauca and Magdalena inter-Andean valleys, plateaus and depressions
of the eastern range, Guajira peninsula, as well as Cesar, San Jorge, Sinú, Atrato and
Magdalena-Cauca river plains. In Orinoquian, Amazonian and Pacific regions, its
presence is negligible [57].
In Valle del Cauca department (inter-Andean zone), it is estimated that there are
around 20,000 ha with certain risk of salinity or sodicity, and 40,000 ha affected by
sodium saturation levels greater than 15% [59]. Here, the progressive advance of
sodicity in agricultural soils is related to inadequate irrigation practices and a lack of
monitoring of water quality parameters (such as bicarbonate content), which pro-
duces a transfer of this anion to the soil [60]. It is also mentioned that the drainage
management in sugarcane crops (predominant in the department) has been
neglected, which prevents the correct washing of salt excess in the soil [61]. In
contrast, in the irrigation district of Alto Chicamocha (Andean region), salinization
of the soils has been ascribed to emanations of thermal waters with content of
sulfates, sodium, and calcium chlorides. Weathering of lake-type sediments depos-
ited in ancient geological times, the high water table, and the salinity of the irrigation
water also contribute to salinization, evinced in more than 50% of the district soils
(4,148 ha) having mild to moderate salinity levels [62].
The Guajira peninsula, in the northernmost part of the country, is the region with
most prevalence of salinity due to its arid condition, in more than 80% of its
extension. The Urabá gulf coast, the greatest producer of exportation bananas in
Colombia, despite having abundant rainfall, suffers from salinization of its soils, as a
consequence of the strong fluctuations in the water table [63]. In Cesar River Valley,
as well as on the eastern bank of the Magdalena River and around the Santa Marta
Great Marsh, salinity is also an acknowledged problem, that can seriously affect the
productivity of oil palm crops that are the agro-industrial engine of the region [9].
It is of general acceptance that addition of manure, compost, and biosolids is a
good reclamation practice for saline soils, which improves physicochemical param-
eters such as electric conductivity, pH, and aggregate stability [64–66]. This practice
was evaluated in a field experiment at Valle del Cauca, showing that the application
of three biowastes (pig manure, composted biosolids, and vinasse) reduced the
exchangeable sodium percentage and the electrical conductivity, preventing the
Agricultural Soil Degradation in Colombia 193

loss of soil productivity in the short and long term [67]. Other experiences also
mentioned the use of cattle manure, molasses, and vinasses for sodic soil reclamation
in the Caribbean region; however, the use of inorganic amendments (gypsum and
sulfuric acid) exhibited better results in relation to pH reduction and sodium dis-
placement in the soil [68]. More recently, the use of low rank coal treated with
solubilizing bacteria that release humified organic matter has proved to be a feasible
alternative for saline-sodic soil reclamation in Colombia [69].

4.1.7 Macro and Microfauna Loss

Recent studies in Colombia report the characteristics and population dynamics of the
soil macrofauna, comparing the cover types, management and altitudinal gradient
[40, 70–72]. This is a recognition of the importance of soil biological component as
an indicator of quality. For instance, the migratory impediment of the earthworm
Martiodrilus carimaguensis was associated with soil degradation, notable by an
increase in compaction and aluminum saturation, and with a decrease in SOC
content and plant biomass. This favored the appearance of opportunistic weeds,
showing that the loss of worms can alter the agroecosystem of the Orinoquian
plains [73].
In conventional grazing paddocks, with very little rotation and high permanence
of cattle, a notable decrease was detected in the populations of adults and larvae of
coleopterans, hemipterans and earthworms, compared to those with intensive rota-
tion. The latter was associated with poor soil quality indicators, in particular bulk
density, porosity, and SOC [40]. Intensive rotation of cattle seems to favor the
abundance of macrofauna, particularly that of worms and beetles, which improve
the structure of the soil by bioturbation. Other authors mention that earthworms seem
to increase their population with agricultural intensification, being found in greater
abundance in grasslands and silvopastoral systems than in forests [71]. However,
this can be in part attributed to the presence of Pontoscolex corethrurus, an invasive
earthworm species associated with degraded landscapes in Amazonian region [71].
In coffee farms of Huila department (Andean region), it was evidenced that the
abundance of macrofauna in soils was mainly concentrated in ants, worms, beetles
and termites. More importantly, significant correlations between the presence of
these taxa and physicochemical properties of the soil were reported, such as sand
percentage, CEC, phosphorus content, and pH [72]. This shows that the alteration of
the physicochemical properties due to inadequate management practices can have a
negative impact on the edaphic macrofauna. After transition from coffee plantations
into pastures, a noticeable reduction in biological diversity was verified, in short
period from 2 to 3 years [29].
Macrofauna groups associated with soil litter, such as decomposers (Collembola
and Diplopoda) and predators (Araneae and Chilopoda), were mainly affected by
the intensification of land use. A strong relationship was found between springtail
density and soil moisture, suggesting that changes in its density are linked to
194 M. Quintero-Angel and D. I. Ospina-Salazar

variations in the microclimate of each type of land use, being more abundant in
cocoa production agroecosystems [71].
In plantain farms situated over 1,200 m.a.s.l. of the Andean region, a distribution
of soil macrofauna is reported, as follows: Oligochaeta (48.6%), Coleoptera
(19.1%), Myriapoda (14.5%), Blattodea (5.5%), and Hemiptera (2.6%) [70]. The
mean abundance of macrofauna was 249.9 ind. m-2 among all farms analyzed. It
highlights the relative importance of earthworms and beetles in the soil; however, the
more intensive the cultivation, the higher the prevalence of exotic species such as
Pontoscolex corethrurus [70].

4.1.8 Contamination

Soils are contaminated naturally or by human action, mostly by pesticides and heavy
metals. The apparent consumption of pesticides in Colombia is 13.5 kg ha-1
between 1990 and 2018, much higher than the world average (2.63 kg ha-1). This
consumption place Colombia on seventh place in the world with the highest con-
sumption of pesticides in its agricultural land, higher than several countries with
more intensive agricultural systems [74]. Other sources place the country in the 17th
place on agrochemical use by hectare with 9.9 kg ha-1 [75]. According to World
Bank, agrochemical consumption in Colombia is 499.4 kg ha-1 [76], much higher
than average value for Latin America (106.9 kg ha-1). However, the last two sources
do not make a distinction between fertilizers and pesticides.
Various sources seem to support the trend of abuse of agrochemicals in Colom-
bia. The National Department of Statistics of Colombia (DANE), estimated the
average annual consumption of fertilizers between 2000 and 2015 at
703.61 kg ha-1 [77]. This value included the consumption of both chemical and
organic fertilizers. The latter may partly explain this high value, since chicken
manure is a commonly used input in national agriculture. Regarding the consump-
tion of pesticides, DANE reported an annual average consumption of 10.19 kg ha-1
in the same period, which is close to the 13.5 kg ha-1 reported in FAOSTAT [74].
Two aspects that can determine the high consumption o agrochemicals in Colom-
bia are (1) reduction of agricultural land use since 2011, due to a change in DANE
methodology which impact apparent consumption calculations of agrochemicals,
and (2) high consumption of herbicides between 1990 and 2018 which was esti-
mated for Colombia in 0.49 kg ha-1, three times higher than world average
(0.16 kg ha-1) [74] (Table 3). This can be related to a high annual consumption of
glyphosate herbicide for illegal crop eradication (mainly coca) by Colombian gov-
ernment. Although since 2016, aspersion with glyphosate for illicit crop eradication
is suspended by ordinance of Constitutional Court, eradication and manual aspersion
continue. By 2015, the last year with aerial aspersion, and area of 36,494 ha in coca
was sprayed [78], at average doses of 4.99 kg of active ingredient per hectare [79].
Other studies evidence negative impact of agrochemicals on soils and ecosys-
tems. Between 1960 and 1970, intensive cotton cultivation with pesticides and
fertilizers, and excessive tillage caused a degradation of thousands of hectares
Agricultural Soil Degradation in Colombia 195

Table 3 Relative consumption of herbicides in Colombia 1990–2018


Unit World Colombia
Herbicide’s consumption 1990–2018 Ton 26,952,182 721,307
Annual average 1990–2018 Ton 929,386 24,873
Area in pastures km2 30,722,240 407,490
Area in crops km2 27,479,430 101,844
Total farming area km2 58,201,670 509,333
Total farming area in hectares ha 5,820,167,047 50,933,334
Relative consumption of herbicide per farming area kg ha-1 0.16 0.49
between 1990 and 2018
Own elaboration with data from FAOSTAT [74]

especially in Cesar river valley [54]. A study from FAO and Health Ministry
revealed that the most affected ecosystems by the use of pesticides are: Ariari,
Caquetá and San Jorge rivers due to rice; Zapatosa marsh by palm oil; Cauca river
by sugar cane; high plateau of Cundinamarca and Boyacá by flowers, potato and
vegetables; Meta, Saldaña and Coello rivers by rice and cotton; and Santa Marta
Great Marsh by banana, palm oil and rice [80]. Additionally, illicit crops (mainly
coca) strongly impact soils and water of sensible ecosystems in the Andean, Ama-
zonian and Pacific regions as they use strong chemical reagents for its processing
(sulfuric acid, hydrochloric acid, potassium permanganate, ketone among
others) [81].
The use of agrochemicals is culturally internalized among rural population,
according to the results of the Third National Agricultural Census of 2014. Chemical
control of pests, diseases and weeds is preferred over other less polluting practices,
as well as the application of chemical fertilizers for soil improvement (Figs. 5 and 6).
However, some authors report very low detection values for pesticide residues in
soils, as these residues are degraded by photolysis during the dry season, and then
leached during the rainy season [82].
In terms of contamination by heavy metals, mercury is by far the most
concerning, impacting the soils of Caribbean, Pacific and some watersheds in the
Andean and Amazonian regions. More than 80 municipalities and 17 departments
face the consequences of mercury contamination, used by artisanal gold extraction
from alluviums. It was estimated that in 2009, 345,570 kg of mercury were released
into the environment, distributed as 31,260 kg in water, 74,420 kg in air, and
151,650 kg in soil [83].
Cadmium is another heavy metal of great concern, especially in regions of cocoa
cultivation, which may reach high soil concentrations due to weathering of the parent
material, by addition of phosphate fertilizers, or by residual deposition of mining and
electronic industry [84]. Cocoa is a cadmium bio-accumulator species and is culti-
vated in rainy regions on acid soils where this metal is more bioavailable. Since there
is a real risk of cadmium for this crop, the need to implement hyper-accumulating
species and/or soil amendments to divert this heavy metal from cocoa trees is
rising [84].
196 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 5 Preferences for pest control in agricultural production units. Based on information from
Departamento Nacional de Estadística-DANE [77]

Fig. 6 Preferences for soil improvement in agricultural production units. Based on information
from Departamento Nacional de Estadística-DANE [77]
Agricultural Soil Degradation in Colombia 197

4.1.9 Nutrient Loss

Natural fertility of Oxisols and Ultisols, predominant in low and humid regions, is
very low per se due to a fast mineralization of organic matter and high weathering of
the solid fraction. However, further nutrient loss due to management and cover type
in natural fertile soils in the Andean zone is attributed to erosion and lixiviation.
In general, over-fertilization in Colombia is widely present, therefore, soil nutri-
ent loss by crop extraction is not a generalized problem. For instance, inputs and
outputs balance of N, P, and K in sugar cane soils has been of positive sign (inputs >
outputs) since 1940 because of fertilizer over application, especially nitrogen.
However, the magnitude of this “unbalance” has been reducing towards the present
time due to better soil management practices, which leads to an optimal nutrient use
through incorporation of crop residues [85]. Nutrient losses through leaching in
Colombian coffee soils depend on its mineralogy, which is highly variable in the
Andean region. Leaching increases with excess water, being always higher for N and
K than for P, given the tendency of this element to fixate in the mineral fraction or to
form insoluble salts [86].
Erosion associated with plant cover type and soil management is the most
important factor of nutrient loss. In hillsides, bare soils exhibit an elevated depletion
of Ca, Mg, N, and K by water erosion during rainy season [14]. In soils with transient
and permanent crops, losses of N, P, K, Ca, and Mg in runoff water were signifi-
cantly superior to pasture and forest soils [15]. Similar results showed a soil loss rate
almost ten times higher in potato cropland than in pastures, which concomitantly
caused losses of N and P by 41.8 and 33.3% [21]. Unexpectedly, a greater loss of N
and P in runoff was registered in potato crops under conservation tillage, mostly due
to greater mineralization of the organic matter from plant biomass incorporated as
green manure [22].

4.2 Soil Degradation Extension in Natural Regions


of Colombia

Although there is no consensus between consulted experts about different soil


degradation processes at regional level, it was possible to identified some patterns
(Fig. 7). Of 81 experts consulted, 61 had experience in Andean region, 22 in
Caribbean region, 22 in Orinoquian region, 21 in Pacific region, 9 in Amazonian
region, and 2 in Insular region.
For the Andean region, most of 61 experts considered that erosion and micro and
macrofauna reduction are very extended soil degradation processes. As well, organic
matter loss is considered as a very extended process by 25 experts, however, other
25 considered it as an extended process. Compaction, nutrient loss, acidification, and
contamination are classified mainly as extended processes and desertification, sali-
nization and sodicity as less extended degradation processes (Fig. 7).
198 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 7 Extension of soil degradation process by regions in Colombia per expert criteria in each
region
Agricultural Soil Degradation in Colombia 199

According to most of the 22 experts in the Caribbean region, in this zone organic
matter loss, erosion, compaction, and micro and macrofauna reduction and nutrient
loss are very extended degradation processes. Desertification is considered as a very
extended process by nine experts, however, other nine considered an extended
process. Salinization and sodicity, acidification, as well as contamination are classi-
fied mainly as extended processes (Fig. 7).
For the Orinoquian region, most of the 22 experts considered compaction,
organic matter loss, and acidification are very extended soil degradation processes.
Erosion, nutrient loss and micro and macrofauna reduction were mainly classified as
extended processes, and desertification as a less extended process. Contamination is
considered as an extended process by six experts and less extended by another six.
Also, salinization and sodicity are not considered relevant soil degradation processes
in this region (Fig. 7).
Most of the 21 experts in the Pacific region, consider erosion, acidification,
nutrient loss, organic matter loss, and micro and macrofauna reduction as very
extended soil degradation processes. Contamination is considered mainly as an
extended process and desertification and compaction as an extended process by six
experts and less extended by another six (Fig. 7).
Most of the nine experts in the Amazonian region considered that compaction,
organic matter loss, acidification and micro and macrofauna reduction are very
extended soil degradation processes. Erosion is classified mainly as an extended
process, while nutrient loss is considered as an extended process by four experts and
less extended by another four. Nevertheless, desertification, salinization and
sodicity, and contamination are not considered as relevant soil degradation processes
in this region (Fig. 7).
In the Insular region, the consensus between the two experts indicates that
compaction, erosion and salinization and sodicity are very extended soil degradation
processes. Erosion is classified as a very extended process by one expert and as a less
extended process by the other (Fig. 7).

4.3 Main Causes and Consequences of Soil Degradation


in Colombia

Although there are contrary opinions, most of the 81 experts consulted agreed that
the main causes of soil degradation in Colombia are: (1) unappropriated planning
and land use planning without considering soil properties (79.01%, 64 experts);
(2) intensive farming practices under technological packages which degrade soil
(77.77%, 63 experts); (3) unawareness of soil ecosystem services, functions and its
importance (77.77%, 63 experts); (4) agricultural frontier expansion by economic
system interests and pressure (76.54%, 62 experts); (5) lack of political will,
disarticulation and state institutionalism weakness (75.30%, 61 experts), (6) lack
of norms, instruments, or its application to sustainable soil management (69.13%,
200 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 8 Expert opinions on soil degradation cause in Colombia

56 experts); (7) knowledge gaps such as few alternatives and technological packages
oriented to soil recovering, restoration and rehabilitation (67.90%, 55 experts);
(8) limited soil quality monitoring (65.43%, 53 experts) (Fig. 8).
Activities producing higher soil degradation in the Andean region correspond to
cattle overgrazing and deforestation for 50 (82%) and 47 (77%) experts, respec-
tively. Crops with no covers and/ or in the hillside are referred by 44 (72.1%)
experts, as well as excessive pesticide and fertilizer application by 39 (63.9%)
experts. Mechanized farming is recognized as a cause of soil degradation by
37 (60.7%) experts, while vegetation burning is recognized by 26 (42.6%) experts,
water quality and quantity in irrigation by 20 (32.8%), as well as unappropriated
waste management by 15 (24.6%) experts (Fig. 9). Some experts highlighted other
activities that favor soil degradation in the Andean region: unappropriated runoff
water management, road construction, soil sealing by urbanization and cropping
systems that affect soil structure such as tuberculous.
In the Caribbean region experts agreed that the main soil degradation activities
correspond to cattle overgrazing and deforestation by 18 (81.8%) and 16 (72.7%)
experts, respectively. Additionally, excessive pesticide and fertilizer application is
referred by 14 (63.6%) experts, water quality and quantity in irrigation by
12 (54.5%) and vegetation burning by 10 (45.5%) experts. Mechanized agriculture
tillage is recognized as soil degradation cause by 8 (36.4%) experts, inadequate
waste management by 7 (31.8%), while crops with no covers and/or in the hillsides
are referred by 4 (18.2%) experts (Fig. 9). Some experts highlighted other activities
that favor soil degradation in this region: overgrazing, cropping in saline/sodic soils
monoculture farming and the lack of soil cover.
Agricultural Soil Degradation in Colombia 201

Fig. 9 Activities causing greater soil degradation by regions

For the Orinoquian region, experts considered that the higher soil degradation
activities, corresponding to deforestation and overgrazing according to 19 (86.4%)
and 17 (77.3%) experts, respectively. Mechanized agriculture tillage is recognized as
a soil degradation cause by 15 (68.2%), vegetation burning by 14 (63.6%) and
excessive pesticide and fertilizer application is referred by 13 (59.1%) experts. Also,
two experts (9.1%) recognized crops with no covers and/or hillside cultivation, and
water quality and quantity in irrigation (Fig. 9), as main soil degradation causes.
Some experts highlighted other land degradation causes such as inadequate crop
selection for acidic soils, inadequate tillage, mechanization and agricultural intensi-
fication of savannas.
The Pacific region experts referred that the main activities causing soil degrada-
tion in this region are deforestation for 19 (90.5%) experts, crops with no covers
and/or hillside cultivation by 17 (80.1%), and vegetation burning by 14 (66.7%)
experts. Cattle overgrazing is referred by 13 (61.9%) experts, excessive pesticide and
fertilizer application by 12 (57.4%), mechanized agriculture tillage by 10 (47.6%),
water quality and quantity in irrigation by 8 (38.1%),3 and inadequate waste man-
agement by 7 (33.3%) experts (Fig. 9). Some experts referred to illegal mining as an
activity that degrade soils in this region, additionally, they suggest that agricultural
intensification under monoculture farming and the loss of traditional cropping
systems are other important factors causing soil degradation in this region.
Amazonian regional experts recognized that main soil degradation activities are
deforestation by 9 (100%), vegetation burning by 8 (88.9%) and cattle overgrazing

3
These answers can be associated with the heavy precipitations, as this region does not use
irrigation but drainage is necessary.
202 M. Quintero-Angel and D. I. Ospina-Salazar

by 6 (66.7%) experts. Also, two experts (22.2%) recognized crops with no covers
and/or hillside cultivation, excessive pesticide and fertilizer application and mech-
anized farming as other degradation causes. Water quality and quantity in irrigation
are referred by one expert (11.1%) and inadequate waste management by none
(Fig. 9). A couple of experts also mention mining as an important factor on soil
degradation in this region.
For the Insular region two experts (100%) agreed that main causes of soil
degradation are deforestation and vegetation burning (Fig. 9). Also, one expert
mentions the use of saline water for irrigation as a problematic practice for soil
degradation in the region.
Among problematics or consequences derived from soil and land degradation in
Colombia, of 81 experts consulted, 73 (90.1%) experts referred alteration and
ecosystems disequilibrium, 64 (79%) increase of rural poverty, 61 (75.3%) decrease
of farming productivity and 60 (74.1%) to the food insecurity. Also, 59 (72.8%)
experts referred to the increase of agricultural frontier by land use change to new
productive lands, 56 (69.1%) increase of sedimentation and water sources turbidity,
and 43 (53.1%) loss of livelihoods. Additionally, some experts suggested other
consequences such as land use conflicts and violence; rural to urban migration, an
increase of landslides emergencies and disasters; increase in CO2 emissions; low
agricultural sustainability and loss of biodiversity, particularly the one in soils which
has been little studied.

4.4 Conflicts by Soil Use in Colombia

Owing to the fact that Colombia is one of the countries that concentrates a large part
of the arable land without exploitation and where the agricultural frontier could be
expanded [87], there is a well-documented conflict between the vocation of soils and
their current use. An excess of cattle ranching and underutilization of soils with
agricultural and forestry vocation are highlighted. In Colombia, 22 million hectares
are suitable for agriculture, 4 million for agroforestry and 15 million for livestock.
However, only 8.5 million hectares are used for agriculture and more than 34 million
for livestock [8].
In the context of the agricultural frontier expansion, deforestation in the Amazon
region has increased significantly, particularly since the 2016s signing of the peace
treaty and the disarmament of Fuerzas Armadas Revolucionarias de Colombia
(FARC), the largest illegal armed group in the country. This deforestation is closely
linked to land grabbing and speculation, and is sustained thanks to extensive cattle
ranching that convert forests into pastures; on many occasions with credits from
public entities, which do not verify the location of the farms, or the vocation of the
land or the management practices used [88]. Contrary to what happens in other South
America and South East Asia countries, soybean and oil palm crop expansion are not
strongly correlated with deforestation [89, 90].
Agricultural Soil Degradation in Colombia 203

Table 4 Soil cover and use in conflict


Present area % of Potential area % of
(millions of national (millions of potential
Use type hectares) territorya hectares) use aptitudeb % in conflict
Natural 63.2 56.7% 64.2 56.2% 0.5%
forest
Crops 8.5 7.6% 22.0 19.3% 11.68% by sub
utilization
Pastures 34.4 30.8% 15.1 13.3% 17.5% by use on not
for cattle suited areas
ranching (overutilization)
Non- 2.5 2.2% N.A. N.A. N.A.
farming
Other 2.8 2.5% N.A. N.A. N.A.
Source: Data from Departamento Nacional de Estadística-DANE [77] and Organización de las
Naciones Unidas para la Alimentación y la Agricultura [91]
a
Third Agricultural National Census 2014
b
FAO

Data from last Agricultural National Census (2014) showed that 56.7% of the
rural area is covered in wild forests, 7.6% in crops, 30.8% in pastures and 4.7% in
other non-agricultural uses (Table 4) [77]. Comparing these figures with the world
average from 2018, where 18.45% of the area was dedicated to crops and 20.63% to
pastures, the trend is confirmed: Colombia dedicates an excess of land to pastures,
while under-utilizing land suitable for crops [74].
According to the potential area with aptitude for forests and the census data,
Colombia has a cover according to its potential. However, its distribution is highly
uneven among natural regions. Most (45.1%) of the natural forests are concentrated
in four departments of the Amazon region, while in some departments of the Andean
and especially at Caribbean region, these land cover has disappeared or are well
below the national average (56.7%); e.g., in Caldas and Sucre forests cover 12.0% of
the area. Similarly, the highest percentage of the area for agricultural use is concen-
trated in four small departments in the Andean region and one in the Orinoquian.
The excess of land under pastures generates a conflict over land use, since it is
present in about 30% of the country, but ideally should only be implemented in
13.3% of the territory because of different geomorphological and environmental
restrictions (high slope, low fertility, scarce or uneven precipitation). Meanwhile, the
agricultural panorama is the opposite: 7.6% of the national territory has crops, while
the potential area is 19.3% [92] (Fig. 10). Therefore, areas under pastures should be
converted into cropland, since there is no margin to expand the agricultural frontier
at the expense of natural forests. Purposely, conversion of pastures into oil palm
crops in the Orinoquian region has been regarded as an economically viable alter-
native and also neutral in carbon emissions [89].
In established agricultural regions, the spread of new croplands can come from
abandoned agricultural areas or by converting from pastures. However, in
204 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 10 Soil use vocation in Colombia. Based on data from Instituto Geográfico Agustín Codazzi
[8]

agricultural frontier areas, the tendency to convert wild areas into new pastures
would continue, counteracting the previous process. The release of pasture land for
crops and forestry would reach about 15 million hectares to achieve a proper
distribution of land use in Colombia [25]. Due to the current narrow benefits of
cattle ranching and inefficient land use, economic incentives can be applied to
convert pastures with intensive CO2 emissions to forest or crops with higher benefits
such as oil palm, sugar cane or coffee [93].
A relevant aspect in land use change and land degradation in Colombia has been
internal conflict between the state and illegal armed groups for over five decades.
Presence and control of these actors, especially guerrillas, in several protected areas
resulted in deforestation discourage. After peace agreement signing in 2016, it was
evident a dramatic increase in deforestation where guerillas were present, mainly in
national parks in the Amazon region and the Catatumbo River watershed. This is a
symptom of state absence in those areas and disconnection with communities
inhabiting those regions [94].
State weakness to control territory, violence and illegal activities also have been
evident in other regions. Although in Colombia deforestation rate in indigenous
communities and afrodescendants territories is up to two times lower than in other
non-ethnic territories [95], these communities faced and try to resist violence of
illegal groups and its pressures towards mechanized mining and drug dealing [96]. In
ethnic territories less deforestation is explained by environment-culture relationship,
traditional knowledge, as well as laws restricting soil use and recognize collective
territorial rights, etc. [95].
Agricultural Soil Degradation in Colombia 205

4.5 Viable Agricultural Practices to Reduce Soil and Land


Degradation

The soil conservation practices commonly used in Colombia’s rural population


correspond mainly to vegetation rewilding, minimum tillage, and direct seeding
[77] (Fig. 11). However, expert consultation showed a higher diversity of actions to
reduce soil degradation. For the Andean region, 51 (83.6%) experts recommended
silvopastoral systems, 45 polyculture production system and live barriers implemen-
tation (73.8%), 45 cover crops or crop residues incorporation into soil (73.8%),
40 sowing in contour lines (65.6%), 38 rotational grazing (62.3%), 37 crop rotation
(60.7%), 33 zero tillage (54.1%), 23 precision and site specific-agriculture (37.7%),
and 11 experts intensive agriculture in greenhouses (18%). Some experts also
suggested a reorientation of farming practices towards other alternative sustainable
frameworks such agroecology, agroforestry (silvicultural, agrosilvopastoral, and
silvopastoralism systems) or conservation agriculture principles; traditional agricul-
tural practices rescued, for example, the wachado [97]; and soil monitoring in
farming facilities (Fig. 12).
For the Caribbean region, 17 (77.3%) experts recommended cover crops or crop
residues incorporation into soil, 14 (63.6%) rotational grazing and silvopastoral
systems, 13 precision and site specific-agriculture (59.1%), and 12 (54.5%) zero
tillage (and crops rotation). Also ten experts (45.5%) recommended polyculture
production system and living barriers, sowing in contour lines, while only one expert
(4.5%) consider intensive agriculture viable in greenhouses. An expert also showed

Fig. 11 Adherence to and type of soil conservation practices in agricultural units. Based on
information from Departamento Nacional de Estadística-DANE [77]
206 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 12 Farming practices to reduce land degradation

that water quality control in irrigation can contribute to reducing soil degradation
(Fig. 12).
In the Orinoquian region, 20 experts recommended silvopastoral system (90.9%),
17 cover crops or crop residues incorporation into soil (77.3%), 16 rotational grazing
(72.7%), 15 (68.2%) crop rotation, as well as precision and site specific-agriculture.
Also 12 experts recommended zero tillage (54.5%), 11 polyculture production
system and living barriers (50%), and 6 (27.3%) sowing in contour lines, while
intensive agriculture in greenhouses is not considered or recommended by experts.
Additionally, experts suggested other options like productive systems implementa-
tion based on agroforestry science and adequate use of efficient irrigation systems.
However, one expert mentioned that the savanna ecosystem should not be consid-
ered for agriculture intensification (Fig. 12).
In the Pacific region, in order to reduce soil degradation, 16 (76.2%) experts
recommended silvopastoral systems, polyculture production system and living bar-
riers implementation. Also, 15 experts (71.4%) suggested crop rotation, cover crops
and crop residues incorporation into soil, and precision and site specific-agriculture.
Other 13 experts recommended rotational grazing (61.9%), 11 sowing in contour
lines (52.3%), 9 zero tillage (42.9%) and 6 intensive agriculture in greenhouses
(28.6%). Additionally, experts referred to other alternatives such agroforestry in
multilayer systems (wood, fruits, leguminous and agricultural crops), restoration of
degraded zones with native species, sector adoption of farming practices adapted to
humid and pluvial tropics. Also, one expert highlighted that this region has a limited
Agricultural Soil Degradation in Colombia 207

agricultural vocation and the importance of conservation of ancestral production


systems from afrodescendants and indigenous communities in this territory (Fig. 12).
In the Amazonian region seven experts (77.8%) recommended silvopastoral
systems (and cover crops or crop residues incorporation into soil), six polyculture
production system and living barriers implementation (66.7%) and five rotational
grazing (55.6%). Also, four experts recommended zero tillage (44.5%), three crop
rotation (33.3%), two precision and site specific-agriculture (22.22%), while only
one expert (11.1%) supports sowing in contour lines and no one recommended
intensive agriculture in greenhouses. One expert even highlighted that Amazonian
region does not have any agricultural vocation whatsoever (Fig. 12).
For the insular zone, two experts (100%) recommended silvopastoral systems,
polyculture production system implementation and living barriers and crop rotation.
Also, one expert recommended cover crops or crop residues incorporation into soil,
zero tillage, sowing in contour lines, precision and site specific-agriculture and
intensive agriculture in greenhouses. However, no expert considered rotational
grazing (Fig. 12).

4.6 Research Gaps and Enabling Factors in Colombia


for Soil and Land Sustainability

Concerning soil and land degradation research in Colombia, the Scopus search
showed 172 publications from 1988 to 2021, with 1,954 different keywords,
which were clustered in 12 research themes and plotted in a strategic map (Fig. 13).
Motor themes in soil and land degradation are in the upper-right quadrant
(Fig. 13) and correspond to sediment-yield, soil-pollutant and crop. These themes
are well developed and are important for research field structuration, also “they are
related externally to concepts applicable to other themes that are conceptually
closely related” [12, 150p]. The thematic areas inside sediment-yield field include
coastal environmental change, erosion rate, delta, mangrove, peat-soil, and Holo-
cene. Soil-pollutant theme includes organic-pollutant, metabolite, endosulfan, insec-
ticide, indoleacetic acid, bacterial protein, enzyme, soil-surface, herbicide,
glyphosate and human. Crop theme includes several fields such as agricultural
production, climate change, trade-flow, crop survey, water resources, rural area,
peace-building, armed conflict, biofuel, and microalgae.
Basic and transversal themes are in the lower-right quadrant, these themes “are
important for a research field but are not developed” [12, 151p]. In this quadrant
there are three themes: Soil, Ecosystems, Acinetobacter sp. The thematic area of soil
theme includes soil degradation, degradation pathway, soil organic matter, nutrient,
growth rate, organic compound, carbon isotope, waste-water, and pasture. Ecosys-
tems theme includes high Andean forest, temperate forest, mangrove, wetland, river,
watershed, ecosystem service, bioindicator, gastropods, above-ground biomass, and
structural equation modeling. Acinetobacter sp. topic includes Bacillus sp.,
208 M. Quintero-Angel and D. I. Ospina-Salazar

Fig. 13 Strategic diagram of SciMAT based on citation (H-index)

biomarker, solid waste, animal, intensive silvopastoral systems, tree, anthropogenic


source, tropical-soil, agricultural chemical, and pesticide.
The themes soil-aquifer-treatment, organic acid and lignocellulosic residue at the
upper-left quadrant (Fig. 13), have only marginal importance for the field, as “are
very specialized and peripheral in character” [12, 150p]. The thematic area of the
theme Soil aquifer treatment includes aquifer and tropical environment. Organic acid
includes thematics such as plant, polysaccharide, phosphate, and fertilizer. Ligno-
cellulosic residue included two fields manure and cellulose.
The lower-left quadrant themes (Fig. 13) are both weakly developed and mar-
ginal, “mainly representing either emerging or disappearing themes” [12, 150p]. In
this quadrant there are two emergent topics: (1) tropical pastures which includes
thematics such as savanna, neotropic, perception, phosphorus form and earthworm
cast; and (2) numerical model which includes thematics such as macroinvertebrates
and column experiment. Finally, the tropical forest theme is located at the
intersection of centrality and density axes, it includes the following thematics:
biogeographical region, forest, REED+, environmental factor, protected area, pri-
mate, fire, plant-root, mycorrhizae, soil contamination, oil-spill.
From systematic search and expert consultation, it is evident that knowledge and
research capacities on land and soil degradation are not homogeneous in the country.
The experts’ majority were from the Andean region, where most of the Colombian
population is concentrated, followed by Caribbean and Orinoquian regions, with less
representativity in the Pacific, Amazonian and Insular regions. The latter is also
reflected in science, technology and innovation advances, and the number of scien-
tific publications by regions. However, it is important to highlight that Colombian
soils are not homogeneous and not all natural regions have agricultural vocation, for
example, insular and Amazonian regions. Therefore, it is necessary to address
Agricultural Soil Degradation in Colombia 209

mainly three research gaps and overcome some disabling factors to an adequate soil
management in Colombia. These gaps and disabling factors correspond to priority
topics identified by authors from the systematic search, expert consultation and their
critical analysis.
• Gap 1. Limited scientific knowledge of soils and low incidence of it in sustainable
land use in Colombia
In Colombia, there is limited research for land and soil knowledge, its conser-
vation, recovery, use and sustainable management, as well as low outreach of the
available research. Similarly, at a regional level, there is limited knowledge of
land and soil degradation, the incidence of different productive and extractive
practices in it, and few soil assessment indicators have been implemented.
According to the national policy of sustainable soil management, limited research
related to soil and land problematics, its characteristics and recovery alternatives
had increased its degradation [1].
For the consulted experts, more research is required in several themes at the
national level: The analysis of effects of different productive practices on soil
fertility and degradation by soil orders and its socio-ecological consequences in
ecosystems and Colombian society quality of life; studies of morphologic, phys-
icochemical and biological properties in soil horizons and its relationship with
agricultural production and soil conservation status; knowledge of potential land
use and productive vocation of soil at a regional level; knowledge, monitoring
and tracing of soil degradation and its effects on function and ecosystem services
loss; evaluation of soil covers with native weeds; determination of thresholds of
soil quality indicators; emergent pollutants studies; study on (un)sustainability of
different soil fertility reposition techniques and determination of nutrient absorp-
tion and extraction curves for different crops and regions; climate change impacts
on soil degradation. Additionally, inventories of soil biodiversity are required, its
genetic identification which is extremely poor, and characterizations of soil
ecosystem services.
Although experts identified relevant research themes, it is important to study
land, soils, and their degradation with an interdisciplinary approach that allows to
correlate changes in soil morphological, physicochemical and biological proper-
ties with changes in food systems or land uses and land covers change, and
particularly, the proximate causes and underlying drivers of change [98]. More-
over, it is a challenge to achieve the understanding of different actors such as
peasant, mestizos, colonist, indigenous, afrodescendants, landowners and cattle
ranchers’ communities, which interests can be opposite. Also, it is necessary to
consolidate knowledge data set of raw data and publications about national soils.
It is also important to finish rural cadastral and delimitation of national wasteland,
as well as ecological sensible areas, for example, moorlands (páramos) and
wetlands.
• Gap 2. Technological dependence and limited local innovation for sustainable
soil and land use
210 M. Quintero-Angel and D. I. Ospina-Salazar

One of the science, technology and innovation challenges in Colombia is to


achieve a real positive impact of the knowledge advances in productive sector and
the well-being of society. However, research centers needed to support Colom-
bian companies to improve their productivity and competitivity are missing [99]
and Colombian society continue to be highly inequitable. Particularly, there is a
limited development of own knowledge and technologies for sustainable land and
soil use in the agricultural sector, based on tropical particularities. Therefore, at
the rural level, most of the supplies, equipment, tools and machinery are imported
from developed countries because of a dependence on external technologies
which in most cases had been developed for different conditions than tropical
ones or are not well adapted.
For consulted experts, more research is required in several themes at the
national level such as: The design, development and adaptation of soil manage-
ment practices; conservation and restoration practices affordable and adapted to
natural regions characteristics of Colombia, and its assessment based on tropical
particularities. Likewise, more research is needed for the design and evaluation of
profitable agricultural productive systems based on natural resource, biodiversity
and soil productive capacity conservation. E.g., design and evaluation of agricul-
tural production systems that simultaneously pressure different soil horizons to
maintain soil nutrition. The development of alternative techniques for soil fertility
reposition based on biomass incorporation and techniques to weed, pest and
disease control oriented to reduce herbicides use, e.g., allelopathic management
based on local biodiversity. The development of tools that support planning and
decision-making for sustainable land and soil use at different scales. E.g, infor-
mation systems, monitoring and field measuring tools, and the development of
field methods for diagnostic and management of soil fertility. Additionally, more
research is needed for the rescue and implementation of technologies and ances-
tral systems for soil and water management.
To approach the relevant themes highlighted by experts, it is necessary to
consider alternative paradigms to conventional agriculture based on monocul-
tures, agrochemical applications and agricultural machinery. In this field, agro-
ecology, economic ecology, agroforestry science, among others are promissory
fields and represent an opportunity to find sustainable soil management options.
• Gap 3. Limited valuation of land and soil importance
Actors in the agricultural sector and Colombian society, in general, have
limited knowledge and valuation of land and soil value, its functioning, and
conservation importance. Likewise, many actors are unaware of the slow process
of soil formation, its relationship with biochemical and hydrological cycles,
therefore there is not an awareness of the causes and effects of soil degradation.
According to the national policy for sustainable soil management [1], in the
country exists deficiencies in education, training and awareness of society on
soils, which leads to a poor understanding of it, the functions, and its importance
for the planet’s subsistence.
For consulted experts, it is required to advance knowledge in different themes
at the national level, such as detailed studies of plural soil valuation (not only
Agricultural Soil Degradation in Colombia 211

economic) by regions; identification and ecosystem services valuation of land and


soil; participatory identification and evaluation of main causes and effects of land
and soil degradation; evaluation of soil knowledge of different actors in the
agriculture sector; pedagogic and didactic strategies oriented to promote land
and soil knowledge by different actors and decision-makers in the agricultural
sector, for example, soil classes in rural schools; environmental education to
generate awareness respect to cause and effects of land and soil degradation;
development, application and adaptation of techniques of research-action-partic-
ipation for soil knowledge and assessment; training and support for producers and
rural technicians to generate, storage and usage of useful information for soil
conservation; development and evolution of social appropriation techniques,
technological transfer and rural extension.
Given the identified research gaps, it is necessary to highlight that Colombia has a
limited budget for science, technology and innovation and important differences
between regions, therefore it is necessary to prioritize research, development and
innovation in macro-regional scientific agendas. The latter, having in mind the
regional environmental differences and that a big part of existing research on soils
is done at the local level. As mentioned by one expert: “research institutions don’t do
it at macro-regional level but at the micro-regional level and therefore there is no
impact.” Without a macro-regional and long-term vision, it would not be possible
project continuity, and it is difficult to address the demands or needs of the territories,
the planning of the acquisition of equipment and laboratories, high-level scientific
training in required areas, etc.
For the definition of macro-regional scientific agendas, workshops could be
organized under research-oriented missions. The latter allows to orient science,
technology, and innovation actions towards strategic areas defined by common
agreement between different actors (public, private, third sector) of the sectors of
interest, and thus address critical problems [100]. Therefore, limited resources and
effort could be prioritized and targeted.
Closing the identified research gaps on land and soil degradation will only be
possible with public policy implementation and public and private investment,
which allows to overcoming some disabling factors. Particularly, in public policies
it is required a consolidation of the national policy implementation to sustainable soil
management [1]. This allows to strengthen and integrate the national institutions that
manage land and soil in a dispersed manner and to improve the planning processes of
land and soil use according to its vocation. Likewise, this allows norm and policy
harmonization, favors soil quality monitoring, as well as compliance of norms to
reduce land and soil degradation, etc.
Land and soil quality monitoring is very important, therefore it is required to
advance towards an information system implementation and monitoring systems
with free access to data and specialized information, at detailed and semi-detailed
scales (1:25,000). A policy oriented to generate public data and information on soils,
its productive vocation, and even crops and productive practices recommendations
according to its features, could reduce soil and land degradation. Currently, this type
212 M. Quintero-Angel and D. I. Ospina-Salazar

of information is available in the country for a couple of industrial crops, based on


private investment in its own research and extension centers, so it is not extended to
other crops.
Straighten scientific capacities for sustainable land and soil management also is
associated with reaching adequate mechanisms of public and private funding.
Science inversion in Colombia, only reached 0.74% of national gross domestic
product in 2019 [101]. From this perspective, it is necessary to improve the univer-
sity, enterprise and state partnership to maximize the impact of science, technology
and innovation. Research centers and institutions face the challenge of searching for
flexible funding mechanisms in cooperation with the industry. Also, a transversal
integration of public research institutions and centers is required in order to share
experiences, resources, laboratories, and complement their agendas and contribu-
tions for science and innovation development [99].
Finally, although this chapter presents a critical situation on land and soil
degradation in different Colombian regions, and the challenges are huge to advance
in science, technology and innovation in this topic, the existence of a public policy
[1] is a factor of hope for closing gaps and overcoming disabling factors.

5 Conclusions

The Colombian natural regions present a wide diversity of soil and land, which has
been threatened since decades ago. The country faces significant soil and land
degradation, primarily compaction, erosion, desertification and salinization, that
challenges the socio-economic activities in the territory.
Stopping agricultural soil and land degradation in Colombia, and if possible, to
reverse this process, will only be possible through state policies implementation and
a paradigm transformation of the stakeholders impacting soil and land with their
activities (farmers, cattle ranchers, among others). Particularly, considering the big
impact of cattle ranching and overgrazing in extended areas of the national territory,
a solid intervention from political and technical level is needed, promoting conser-
vation soil use based on tree-crop-cattle systems, reducing its impacts and environ-
mental passives. Also, it should be considered pasture reconversion to forestry
production systems or permanent crops, offsetting crop land deficit.
Regarding research and innovation on land and soil degradation in Colombia,
three gaps or current problematics were identified which demands immediate and
effective actions. These are (1) Limited scientific knowledge of soils and low
incidence of it in sustainable land use in Colombia; (2) Technological dependence
and limited local innovation for sustainable soil and land use; (3) Limited valuation
of land and soil importance.
Closing these research gaps require public policy implementation such as the
national policy for soil sustainable management. Its implementation will allow to
strengthen and integrate national institutions that manage land and soil, improve the
land use planning, among others. Also, public and private investment is required to
Agricultural Soil Degradation in Colombia 213

establish long-term research agendas by macro-regions, which would allow priori-


tizing and directing resources and efforts. Finally, public data generation and
information about land and soils, their productive vocation and even recommended
productive practices according to its features, could be useful to reduce land and soil
degradation in the country.

References

1. Ministerio de Ambiente y Desarrollo Sostenible-MADS (2016) Politica para la gestión


sostenible del suelo. Comunicaciones MADS, República de Colombia
2. Ruiz F, Arango C, Dorado J, Guzmán D (2012) Cambio climático más probable para
Colombia a lo largo del siglo XXI respecto al clima presente. Nota Técnica. IDEAM, Bogotá
3. Sánchez-López R, Gómez-Sánchez C, Palacios-Fernández A et al (2012) Programa nacional
de monitoreo y seguimiento de la degradación de suelos y tierras en Colombia. IDEAM,
Bogotá
4. IDEAM-Instituto de Hidrología Meteorología y Estudios Ambientales (2010) Segunda
Comunicación Nacional ante la Convención Marco de las Naciones Unidas sobre Cambio
Climático. IDEAM, Bogotá
5. Departamento Administrativo Nacional de Estadística-DANE (2018) Censo nacional de
población y vivienda. https://www.dane.gov.co/index.php/estadisticas-por-tema/demografia-
y-poblacion/censo-nacional-de-poblacion-y-vivenda-2018. Accessed 1 May 2021
6. Llinás-Rivera R (2013) Delimitación de las regiones naturales de Colombia, Exposicion.
Sociedad Geográfica de Colombia, Bogotá
7. Hahn-de-Castro LW, Meisel-Roca A, Hahn-De-Castro LW (2018) La desigualdad económica
entre las regiones de Colombia, 1926-2016. Cuad Hist Económica:1–30. https://doi.org/10.
32468/chee.47
8. Instituto Geográfico Agustín Codazzi IGAC (2012) Estudio de los conflictos de uso del
territorio colombiano escala 1:100.000. Imprenta Nacional de Colombia, Bogotá
9. García-Ocampo A (2012) Fertility and soil productivity of Colombian soils under different soil
management practices and several crops. Arch Agron Soil Sci 58:S55–S65. https://doi.org/10.
1080/03650340.2012.700510
10. Smit B, Wandel J (2006) Adaptation, adaptive capacity and vulnerability. Glob Environ Chang
16:282–292. https://doi.org/10.1016/j.gloenvcha.2006.03.008
11. Cobo MJ, López-Herrera AG, Herrera-Viedma E, Herrera F (2012) SciMAT: a new science
mapping analysis software tool. J Am Soc Inf Sci Technol 63:1609–1630. https://doi.org/10.
1002/asi.22688
12. Cobo MJ, López-Herrera AG, Herrera-Viedma E, Herrera F (2011) An approach for detecting,
quantifying, and visualizing the evolution of a research field: a practical application to the
Fuzzy Sets Theory field. J Informet 5:146–166. https://doi.org/10.1016/j.joi.2010.10.002
13. IDEAM-Instituto de Hidrología Meteorología y Estudios Ambientales, UDCA-Universidad de
Ciencias Aplicadas y Ambientales (2015) Síntesis del estudio nacional de la degradación de
suelos por erosión en Colombia – 2015. IDEAM – MADS, Bogotá
14. Rodríguez JA, Sepúlveda IC, Camargo-García JC, Galvis-Quintero JH (2009) Pérdidas de
suelo y nutrientes bajo diferentes coberturas vegetales en la zona Andina de Colombia. Acta
Agronómica 58:160–166
15. Suescún D, Villegas JC, León JD et al (2017) Vegetation cover and rainfall seasonality impact
nutrient loss via runoff and erosion in the Colombian Andes. Reg Environ Chang 17:827–839.
https://doi.org/10.1007/s10113-016-1071-7
214 M. Quintero-Angel and D. I. Ospina-Salazar

16. Labrière N, Locatelli B, Laumonier Y et al (2015) Soil erosion in the humid tropics: a
systematic quantitative review. Agric Ecosyst Environ 203:127–139. https://doi.org/10.
1016/j.agee.2015.01.027
17. Álvarez-Herrera JF, Fernández JP (2009) Evaluación de la erosión de un inceptisol de Tunja
con diferentes coberturas al impacto de lluvias simuladas. Ing e Investig 29:86–91
18. Carretta L, Tarolli P, Cardinali A et al (2021) Evaluation of runoff and soil erosion under
conventional tillage and no-till management: a case study in Northeast Italy. Catena 197:
104972. https://doi.org/10.1016/j.catena.2020.104972
19. McGregor DFM (1980) An investigation of soil erosion in the Colombian rainforest zone.
Catena 7:265–273. https://doi.org/10.1016/S0341-8162(80)80018-X
20. Sonder K, Müller-Sämann K, Hilger T, Leihner D (2002) Erosion control and prediction in
cassava based cropping systems in the southern andean region of Colombia. In: 12th ISCO
conference. International Soil Conservation Organization, Beijing
21. Otero JD, Figueroa A, Muñoz FA, Peña MR (2011) Loss of soil and nutrients by surface runoff
in two agro-ecosystems within an Andean paramo area. Ecol Eng 37:2035–2043. https://doi.
org/10.1016/j.ecoleng.2011.08.001
22. Uribe N, Corzo G, Quintero M et al (2018) Impact of conservation tillage on nitrogen and
phosphorus runoff losses in a potato crop system in Fuquene watershed, Colombia. Agric
Water Manag 209:62–72. https://doi.org/10.1016/j.agwat.2018.07.006
23. Hoyos N (2005) Spatial modeling of soil erosion potential in a tropical watershed of the
Colombian Andes. Catena 63:85–108. https://doi.org/10.1016/j.catena.2005.05.012
24. Suárez de Castro F (1953) Pérdidas de suelo y agua en un cafetal y en un potrero. Boletín
Técnico Fed Nac Cafe 1:1–13
25. Vargas G, León N, Hernández Y (2019) Agricultural socio-economic effects in Colombia due
to degradation of soils. In: Meena R, Kumar S, Bohra J, Jat M (eds) Sustainable management
of soil and environment. Springer, Singapore, pp 289–337
26. Martınez LJ, Zinck JA (2004) Temporal variation of soil compaction and deterioration of soil
quality in pasture areas of Colombian Amazonia. Soil Tillage Res 75:3–18. https://doi.org/10.
1016/j.still.2002.12.001
27. Pinzón A, Amézquita E (1991) Compactación de suelos por el pisoteo de animales en pastoreo
en el piedemonte amazónico de Colombia. Pasturas Trop 13:21–26
28. Polanía-Hincapié KL, Olaya-Montes A, Cherubin MR et al (2021) Soil physical quality
responses to silvopastoral implementation in Colombian Amazon. Geoderma 386:114900.
https://doi.org/10.1016/j.geoderma.2020.114900
29. Sadeghian S, Rivera JM, Gómez ME (1999) Impacto de sistemas de ganadería sobre las
características físicas, químicas y biológicas de suelos en los Andes de Colombia. In:
Agroforestería para la producción animal en América Latina. Organización de las Naciones
Unidas para la Agricultura y la Alimentación, Roma
30. Martínez J, Cajas YS, León JD, Osorio NW (2014) Silvopastoral systems enhance soil quality
in grasslands of Colombia. Appl Environ Soil Sci 2014:359736. https://doi.org/10.1155/2014/
359736
31. Cherubin MR, Chavarro-Bermeo JP, Silva-Olaya AM (2019) Agroforestry systems improve
soil physical quality in northwestern Colombian Amazon. Agrofor Syst 39:1741–1753. https://
doi.org/10.1007/s10457-018-0282-y
32. Madero-Morales E, Peña-Artunduaga ME, Escobar BY, Fernando García L (2012)
Compactación potencial en dos suelos de la parte plana del Valle del Cauca. Acta Agronómica
61:27–31
33. Rodríguez LA, Valencia JJ (2012) Impact of traffic equipment during sugarcane (Saccharum
officinarum) harvest. Rev Bras Eng Agrícola e Ambient 16:1128–1136. https://doi.org/10.
1590/S1415-43662012001000014
34. Rodríguez LA, Valencia JJ, Urbano JA (2012) Soil compaction and tires for harvesting and
transporting sugarcane. J Terramechanics 49:183–189. https://doi.org/10.1016/j.jterra.2012.
04.002
Agricultural Soil Degradation in Colombia 215

35. Torres JS, Villegas F (1995) Differentiation of soil compaction and cane stool damage. In:
XXI congress of the international society of sugar cane technologists. Kasetsart Univ.,
Bangkok, pp 294–305
36. Madero-Morales E, Ramírez-Alzate JA, Albán A et al (2011) Compaction on sugarcane
cultivated soils in the south area of Valle del Cauca, Colombia. Acta Agronómica 60:242–250
37. Ordoñez M-C, Casanova Olaya JF, Galicia L, Figueroa A (2020) Soil carbon dynamics under
pastures in Andean socio-ecosystems of Colombia. Agronomy 10:507. https://doi.org/10.
3390/agronomy10040507
38. León JD, Osorio NW (2014) Role of litter turnover in soil quality in tropical degraded lands of
Colombia. Sci World J 2014:693981. https://doi.org/10.1155/2014/693981
39. Navarrete D, Sitch S, Aragão LEOC, Pedroni L (2016) Conversion from forests to pastures in
the Colombian Amazon leads to contrasting soil carbon dynamics depending on land man-
agement practices. Glob Chang Biol 22:3503–3517. https://doi.org/10.1111/gcb.13266
40. Teutscherová N, Vázquez E, Sotelo M et al (2021) Intensive short-duration rotational grazing
is associated with improved soil quality within one year after establishment in Colombia. Appl
Soil Ecol 159:103835. https://doi.org/10.1016/j.apsoil.2020.103835
41. Instituto Geográfico Agustín Codazzi-IGAC (2014) Manejo de Suelos Colombianos. Imprenta
Nacional de Colombia, Bogotá
42. Rincón-Castillo A, Baquero-Peñuela JE, Flórez-Díaz H (2012) Manejo de la nutrición mineral
en sistemas ganaderos de los Llanos Orientales de Colombia. Corpoica, Villavicencio
43. Sadeghian S (2016) La acidez del suelo, una limitante para la producción de café. Av técnicos
CENICAFÉ 466:1–12
44. Vallejo VE, Gómez MM, Cubillos AM, Roldán F (2011) Effect of land use on the density of
nitrifying and denitrifying bacteria in the Colombian Coffee Region. Agron Colomb 29:455–
464
45. Parnell RA, Burke KJ (1990) Impacts of acid emissions from Nevado del Ruiz volcano,
Colombia, on selected terrestrial and aquatic ecosystems. J Volcanol Geotherm Res 42:69–88.
https://doi.org/10.1016/0377-0273(90)90070-V
46. Castro HE, Gómez MI, Munévar OE, Hernández DM (2006) Diagnóstico y control de la
acidez en suelos sulfatados ácidos en el Distrito de riego del Alto Chicamocha (Boyacá)
mediante pruebas de incubación. Agron Colomb 24:122–130
47. Combatt-Caballero E, Jarma-Orozco A, Paternina-Durango E (2015) Bromatología de
Brachiaria decumbens Stapf y Cynodon nlemfuensis Vanderyst en suelos sulfatados ácidos
en Córdoba, Colombia. Rev Mex ciencias agrícolas 6:1035–1049
48. Sanchez-Torres JD (2012) Acidez de los suelos y su manejo. Boletín Técnico
Cenibanano:1–10
49. Haynes RJ, Hamilton CS (1999) Effects of sugarcane production on soil quality: a synthesis of
world literature. Proc S Afr Sug Technol Ass:45–51
50. Mina-Possú W (2016) Incidencia del cultivo de la caña de azúcar en la fertilidad del suelo y su
efecto en la producción de alimento de la vereda Agua Azul municipio de Villa rica – Cauca.
Universidad de Manizales
51. United Nations (1994) United Nations Convention to Combat Desertification (UNCCD).
https://www.unccd.int/. Accessed 1 May 2021
52. Vargas-Cuervo G, Gómez CE (2003) La desertificación en Colombia y el cambio global. Cuad
Geogr Rev Colomb Geogr 12:121–134
53. IGAC, IDEAM, MAVDT (2010) Protocolo para la identificación y evaluación de los procesos
de degradación de suelos y tierras por desertificación. Instituto Geográfico Agustín Codazzi;
Instituto de Hidrología, Meteorología y Estudios Ambientales y El Ministerio de Ambiente,
Vivienda y Desarrollo Territorial, Bogotá
54. Wagner-Medina GM (2011) Las huellas ambientales del oro blanco - la expansión algodonera
en el Valle del Río Cesar (1950–1980). Universidad de los Andes
55. Rojas-Marín CA, Pérez-Gómez U, Fernández-Méndez F (2019) Dinámica espaciotemporal de
los procesos de desertificación y revegetalización natural en el enclave seco de La Tatacoa,
216 M. Quintero-Angel and D. I. Ospina-Salazar

Colombia/rcdg.v28n1.63130. Cuad Geogr Rev Colomb Geogr 28:134–151. https://doi.org/10.


15446/rcdg.v28n1.63130
56. Contreras D, Voets A, Junghardt J et al (2020) The drivers of child mortality during the
2012–2016 drought in La Guajira, Colombia. Int J Disaster Risk Sci 11:87–104. https://doi.
org/10.1007/s13753-020-00255-0
57. IDEAM-Instituto de Hidrología Meteorología y Estudios Ambientales, UDCA-Universidad de
Ciencias Aplicadas y Ambientales (2017) Mapa nacional de degradación de suelos por
salinización. http://www.ideam.gov.co/documents/24277/69989379/Lanzamiento+mapa
+Salinizacion+FN+OPT.pdf/624515d0-799d-41ef-b1ef-bb7e868680f3. Accessed
20 Apr 2021
58. Bauder TA, Davis JG, Waskom RM (2014) Managing saline soils. Fact sheet no. 0.503, crop
series soil. In: Color. State Univ. Ext. www.ext.colostate.edu/pubs/crops/00503.pdf. Accessed
20 Apr 2021
59. Calero-Aguado A, Romero-Lozada GA, Martínez-Herrera C (2013) La salinidad de los suelos
en el valle del río Cauca. Tecnicaña 31:27–35
60. Ramírez-Alzate JA (2011) Evaluación general de la salinidad y modelación de los riesgos de
salinización en suelos del Valle del Cauca. Universidad Nacional de Colombia, Sede Palmira
61. CENICAÑA (2013) Agua: un recurso que compromete a ingenios y cultivadores.
Cart Inf:12–13
62. Girón-Pinto JD (2019) Evaluación documental de los métodos de restauración de suelos
salinos, con influencia en el distrito de riego Usochicamocha, departamento de Boyacá.
Universidad de la Salle
63. Durango JC, Mercado-Fernández T, Feria-Díaz JJ (2020) Effect of shallow water table on
banana crop (Musa AAA) in the Urabá subregion in Colombia. J Xi’an Univ Archit
Technol XII:3248–3255
64. Liang Y, Si J, Nikolic M et al (2005) Organic manure stimulates biological activity and barley
growth in soil subject to secondary salinization. Soil Biol Biochem 37:1185–1195. https://doi.
org/10.1016/j.soilbio.2004.11.017
65. Wang L, Sun X, Li S et al (2014) Application of organic amendments to a coastal saline soil in
North China: effects on soil physical and chemical properties and tree growth. PLoS One 9:1–
9. https://doi.org/10.1371/journal.pone.0089185
66. Zhang T, Wang T, Liu KS et al (2015) Effects of different amendments for the reclamation of
coastal saline soil on soil nutrient dynamics and electrical conductivity responses. Agric Water
Manag 159:115–122. https://doi.org/10.1016/j.agwat.2015.06.002
67. Gutiérrez MA, Zúñiga O, Ospina-Salazar DI (2016) Effect of three biowastes on the produc-
tivity potential of a sodic soil. Agron Colomb 34:250–259. https://doi.org/10.15446/agron.
colomb.v34n2.55044
68. Duran R, García J, Amaya R (2000) Evaluación de varias enmiendas para la corrección de
suelos sódicos en el valle del Cesar. Suelos Ecuatoriales 30:21–28
69. Cubillos-Hinojosa JG, Valero NO, Melgarejo LM (2015) Assessment of a low rank coal
inoculated with coal solubilizing bacteria as an organic amendment for a saline-sodic soil.
Chem Biol Technol Agric 2:21. https://doi.org/10.1186/s40538-015-0048-y
70. Rojas-Múnera DM, Feijoo-Martínez A, Molina-Rico LJ et al (2021) Differential impact of
altitude and a plantain cultivation system on soil macroinvertebrates in the Colombian Coffee
Region. Appl Soil Ecol 164:103931. https://doi.org/10.1016/j.apsoil.2021.103931
71. Rodríguez-Suárez L, Suárez-Salazar JC, Casanoves F et al (2021) Cacao agroforestry systems
improve soil fertility: comparison of soil properties between forest, cacao agroforestry sys-
tems, and pasture in the Colombian Amazon. Agric Ecosyst Environ 314:107349. https://doi.
org/10.1016/j.agee.2021.107349
72. Machado-Cuéllar L, Rodríguez-Suárez L, Murcia-Torrejano V et al (2020) Macrofauna del
suelo y condiciones edafoclimáticas en un gradiente altitudinal de zonas cafeteras, Huila,
Colombia. Rev Biol Trop 69:102–112. https://doi.org/10.15517/rbt.v69i1.42955
Agricultural Soil Degradation in Colombia 217

73. Decaëns T, Jiménez JJ, Lavelle P (1999) Effect of exclusion of the anecic earthworm
Martiodrils carimaguensis Jiménez and Moreno on soil properties and plant growth in grass-
lands of the eastern plains of Colombia. Pedobiologia (Jena) 43:835–841
74. FAOSTAT (2021) Food and agriculture data. http://www.fao.org/faostat/en/#data/EP/
visualize
75. Worldometers (2021) Pesticide use by country. https://www.worldometers.info/food-
agriculture/pesticides-by-country/. Accessed 21 Jan 2021
76. Uribe Pérez S (2013) Agroquímicos envenenan suelos en Colombia. UN Periódico:6–7
77. Departamento Nacional de Estadística-DANE (2016) Tercer Censo Nacional Agropecuario
2014, vol 2. DANE, Bogotá
78. Observatorio de drogas de Colombia (2021) Estadísticas Nacionales. http://www.odc.gov.co/
sidco/perfiles/estadisticas-nacionales
79. Solomon KR, Anadón A, Cerdeira AL et al (2005) Estudio de los efectos del programa de
erradicación de cultivos ilícitos mediante la aspersión aérea con el herbicida glifosato (PECIG)
y de los cultivos ilícitos en la salud humana y en el medio ambiente. Informe preparado para la
Comisión Interamericana, Washington
80. Instituto Geográfico Agustín Codazzi-IGAC (2016) Las 6 “plagas” que causan la muerte de los
suelos colombianos
81. Comunidad Andina (2013) Manual de Sustancias Químicas usadas en el Procesamiento de
Drogas Ilícitas, Lima
82. Martínez-Mera EA, Torregroza-Espinoza AC, Crissien-Borrero TJ et al (2019) Evaluation of
contaminants in agricultural soils in an irrigation district in Colombia. Heliyon 5. https://doi.
org/10.1016/j.heliyon.2019.e02217
83. Martínez-Sepúlveda JA (2018) Contaminación y remediación de suelos en Colombia:
aplicación a la minería de oro. Universidad EAN, Bogotá
84. Benavides-Bolaños JA (2019) Phytoextractor species and dolomitic lime as strategies to
manage Cd in cacao (Theobroma cacao) and spinach (Spinacia oleracea). The Pennsylvania
State University
85. Delgadillo-Vargas O, Garcia-Ruiz R, Forero-Álvarez J (2016) Fertilising techniques and
nutrient balances in the agriculture industrialization transition: the case of sugarcane in the
Cauca river valley (Colombia), 1943–2010. Agric Ecosyst Environ 218:150–162. https://doi.
org/10.1016/j.agee.2015.11.003
86. Sadeghian S, González-Osorio H, Arias-Suárez E (2015) Lixiviación de nutrientes en suelos
de la zona cafetera. Prácticas que ayudan a reducirla. Boletín técnico CENICAFÉ 40:1–34
87. Food and Agriculture Organization-FAO (2015) World agriculture: towards 2015/2030.
Summary report. FAO, Rome
88. Rojas T (2021) Amazonia, una selva que arde. El Tiempo
89. Quezada JC, Etter A, Ghazoul J et al (2019) Carbon neutral expansion of oil palm plantations
in the Neotropics. Sci Adv 5. https://doi.org/10.1126/sciadv.aaw4418
90. De Sy V, Herold M, Achard F et al (2015) Land use patterns and related carbon losses
following deforestation in South America. Environ Res Lett 10:1–15
91. Organización de las Naciones Unidas para la Alimentación y la Agricultura-FAO (2018) Guía
de buenas prácticas para la gestión y uso sostenible de los suelos en áreas rurales. FAO y
MADS, Bogotá
92. Nieto-Escalante JA (2019) Colombia, un país con una diversidad de suelos ignorada y
desperdiciada. https://igac.gov.co/es/noticias/colombia-un-pais-con-una-diversidad-de-
suelos-ignorada-y-desperdiciada. Accessed 1 Feb 2021
93. ORBITAS (2021) La carne bovina colombiana. Informe de análisis de riesgos de transición
climática. https://sociedadsostenible.co/wp-content/uploads/2021/04/Orbitas-La-carne-
bovina-colombiana-Informe-de-análisis-de-riesgos-de-transición-climática-2021.pdf.
Accessed 22 Mar 2021
218 M. Quintero-Angel and D. I. Ospina-Salazar

94. Clerici N, Armenteras D, Kareiva P et al (2020) Deforestation in Colombian protected areas


increased during post-conflict periods. Sci Rep 10:4971. https://doi.org/10.1038/s41598-020-
61861-y
95. FAO, FILAC (2021) Los pueblos indígenas y tribales y la gobernanza de los bosques. Una
oportunidad para la acción climática en América Latina y el Caribe. FAO, Santiago
96. Vélez MA, Robalino J, Cardenas JC et al (2020) Is collective titling enough to protect forests?
Evidence from Afro-descendant communities in the Colombian Pacific region. World Dev
128:104837. https://doi.org/10.1016/j.worlddev.2019.104837
97. Volverás-Mambuscay B, Merchancano-Rosero JD, Campo-Quesada JM, López-Rendón JF
(2020) Propiedades físicas del suelo en el sistema de siembra en wachado en Nariño,
Colombia. Agron Mesoam 31:731–748
98. Geist HJ, Lambin EF (2002) Proximate causes and underlying driving forces of tropical
deforestation tropical forests are disappearing as the result of many pressures, both local and
regional, acting in various combinations in different geographical locations. Bioscience 52:
143–150
99. Misión Internacional de Sabios 2019 (2020) Colombia hacia una sociedad del conocimiento.
Reflexiones y Propuestas. Vicepresidencia de la República de Colombia y Ministerio de
Ciencia, Tecnología e Innovación, Bogotá
100. Mazzucato M (2018) Missions: mission-oriented research & innovation in the European
Union. A problem-solving approach to fuel innovation-led growth, Brussels
101. Lucio-Álvarez J, Guevara Rey A, Perea G et al (2020) Indicadores de ciencia y tecnología
Colombia 2019. Observatorio Colombiano de Ciencia y Tecnología, Bogotá
Agricultural Land Degradation in India

Shoba Periasamy and Ramakrishnan S. Shanmugam

Contents
1 Introduction to Agricultural Land Degradation in India . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
2 Policy Framework of India for Agriculture Promotion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
3 Land Degradation Status in India . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4 Land Degradation Status in the Climatic Zones of India . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
5 Land Degradation Processes in India . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
5.1 Soil Erosion and Overgrazing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
5.2 Salinity/Alkalinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
5.3 Waterlogging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
5.4 Agrochemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
5.5 Slash-and-Burn Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
5.6 Microplastics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
5.7 Fluoride . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
6 State-Wise Land Degradation Status . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
7 Mitigation Strategies Adopted in Agro-Climatic Zones of India . . . . . . . . . . . . . . . . . . . . . . . . . . 240
7.1 Himalayan Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
7.2 Indo-Gangetic Plains (IGP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
7.3 Dry and Arid Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
7.4 Coastal Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
7.5 Red and Black Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
7.6 Slash-and-Burn Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
8 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249

Abstract Agriculture plays a vital role in the Indian economy, contributing 19.9%
to the country’s total Gross Domestic Product (GDP) and providing employment to a
large proportion of the country’s population. Being a country known for its

S. Periasamy (✉)
Department of Civil Engineering, SRM Institute of Science and Technology, Chengalpattu,
India
R. S. Shanmugam
Institute of Remote Sensing, Anna University, Chennai, India

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 219
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 219–258, DOI 10.1007/698_2022_913,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 14 December 2022
220 S. Periasamy and R. S. Shanmugam

physiographic diversity with desert, plains, mountains, coastal plains, and peninsular
plateau and ever-increasing population, Indian agricultural land has been prone to a
wide variety of physical and chemical stresses from natural and anthropogenic
channels. In the country, nearly 29% (96.4 Mha) of India’s total geographical area
(328.7 Mha) is subjected to degradation majorly by soil erosion by water
(36.1 Mha), followed by vegetation degradation, referred to as a decline in the
above-ground biomass resulted from deforestation/overgrazing (29.3 Mha), wind
erosion (18.2 Mha), salinity (3.7 Mha), frost shattering (3.3 Mha), and human
interventions that include mining, urbanization, and industrial activities (2.3 Mha),
mass movement or mass wasting that includes all forms of downward movement of
soil and rock under the influence of gravity (0.9 Mha), waterlogging (0.7 Mha), and
others (1.9 Mha). India has focused on reducing degraded areas, particularly from
1996 through the new policy framework with public investments made in the
agricultural domain to achieve a neutral status in land degradation by 2030. How-
ever, while focusing on the existing major degradation processes like water and wind
erosion, salinity, and waterlogging, emerging pollutants like microplastics and
fluoride are given the least priority, which may result in a significant threat to
environmental sustainability and food security in the future. This chapter’s objective
is to highlight the status of soil degradation in India with particular emphasis on the
existing and emerging degradation processes such as soil erosion by wind and water,
overgrazing, salinity, waterlogging, slash-and-burn agriculture, agrochemicals,
microplastics, and fluoride pollution. This chapter also emphasizes the leading
causes and effects of various physical and chemical land degradation processes,
macro- and micro-scale statistics of different degradation processes, and national
interventions in conserving and reclaiming the affected lands.

Keywords Driving forces, Mitigation strategies, National policy framework,


Processes, Soil degradation status

1 Introduction to Agricultural Land Degradation in India

Agriculture contributes about 19.9% to the country’s total GDP and plays a major
role in the Indian economy by contributing a significant role in national output and
providing employment to around 54.6% of the country’s workforce [1, 2]. However,
ensuring environmental sustainability and food security to the population is still
challenging as agricultural development has taken place at the cost of soil degrada-
tion and biomass loss [3, 4]. India is experiencing huge monetary losses from the
agricultural sector as agricultural land degradation has become a severe problem in
the rain-fed and irrigated areas [5]. Land degradation processes are progressing
through various natural and anthropogenic channels. Climate change and
non-climate stresses (growth of population and cattle community) are increasingly
adding pressure on the natural resources and thus result in unsustainable land
Agricultural Land Degradation in India 221

management. Soil physical and chemical characteristics, like texture, compaction,


depth, and mineral composition, can determine soil susceptibility to land degrada-
tion processes. It was projected that nearly 121 Mha of land in India has been
degraded from six decades of green revolution [6] due to intensive agricultural
activities and lack of awareness of emerging pollutants. Despite the implementation
of various agrarian promotion schemes [7], recent reports stated that nearly 64 out of
142 Mha of net sown agricultural land [8] in India was estimated to be severely
degraded [5] due to various physical and chemical stresses, including soil erosion by
wind and water [9], salinity [10, 11], water-logging, overgrazing [5], slash-and-burn
agricultural practice [12], and use of agrochemicals [13], microplastics [14], and
other geogenic pollutants such as fluoride (F) [15, 16] and arsenic (As) [17, 18].
Land degradation and sustainable food production in India are highly influenced
by climate change [19], geography [20], and ever-increasing population [21]. Hence,
maintaining the ecological balance between the present and future population is still
challenging for sustainable land development [22]. Land degradation processes are
closely associated with one another, based on the intensity and duration of climate
and non-climate stresses toward the vulnerable land. The occurrence of one event for
a considerable period may increase the possibility of occurrence of another event.
The awareness of its adverse effect on the environment is vital to plan sustainable
and more resilient solutions for the future. India has committed to increasing its food
production by 45% by 2050 [23] from the decreasing cultivable land (20%) [21] for
the estimated population of 1.6 billion [23]. It necessitates the need to reclaim the
degraded and degrading agricultural land. Being a signatory of the United Nations
Convention on Combating Desertification (UNCCD), India has committed to
achieving a neutral land degradation status by 2030. Hence, designing a sustainable
farming strategy to prevent the agricultural soil from the land degradation processes
is gaining attention by the policymakers in the recent era. This chapter gives the
reader insight into the national policies for agricultural promotion, macro- and
micro-scale perception of India’s land degradation processes, the driving forces,
their contribution to the agricultural loss, and mitigation strategies adopted to
conserve the soil.

2 Policy Framework of India for Agriculture Promotion

In this section, several Indian agricultural policies and their effects on agricultural
growth are emphasized. The evolution of Indian agricultural promotion schemes was
categorized into five phases 1951–1965, 1966–1981, 1981–1991, 1992–2004, and
2005–2013. In the first phase (1951–1965), the industrial promotion strategy was
given high priority with limited focus on agriculture. During this phase, the agricul-
tural area increased without any productivity gains, so the food requirements of the
population had addressed by importing agricultural commodities from the USA
through the PL-480 agreement [24]. Due to the sudden disruption of food supplies
under the PL-480 program, India experienced a severe food shortage in the early
222 S. Periasamy and R. S. Shanmugam

1960s. Achieving food self-sufficiency in a short period became the primary goal of
policy planning [25]. The second phase (1966–1981) was the most remarkable phase
in the history of Indian agriculture with the introduction of Green Revolution or
High-Yielding Variety Seeds (HYV seeds) technology to attain self-sufficiency in
food production [26]. This policy has eradicated large-scale famine by producing
huge quantities of food grains mainly rice and wheat, due to which India has
achieved self-sufficiency in food grain production in the mid-1970s. In the third
phase (1981–1991), the agricultural growth was equalized in all parts of India. The
attention was shifted to the production of oilseeds and pulses as self-sufficiency was
already met in food grain through the green revolution. The Technology Mission on
Oilseeds and Pulses (TMOP) was initiated in 1986 to promote the growth and
increase the production of oilseeds and pulses [27].
During the fourth phase (1992–2004), the macro-economic reforms were
launched to reduce the anti-agricultural bias [28]. Despite a reduction in anti-
agricultural bias, the sharp decline in agricultural production was witnessed in the
second stage of this phase due to decreased capital investment and farm productivity.
This was due to the rapid increase of subsidies on power, fertilizer, and food that led
to the decline in public investment in agriculture, resulting in an adverse effect on
agriculture performance in the 1990s. The country’s GDP dropped to 3.1 in the
1990s compared to 3.5 in the 1980s, resulting in a severe agricultural crisis in the late
1990s. Though the average growth rate of agriculture increased steadily after the
green revolution from 2.1 to 3.5, it came down drastically to 2.2 (equal to the
pre-green revolution phase) during 1997–2004, because of the decline in public
investment in the agricultural sector that resulted in a reduction in the supply of
inputs required for agricultural activity [25]. In the fifth phase (2005–2013), signif-
icant efforts were undertaken by the country in reforming the policy to address these
issues, including the decline in public investment in agriculture, making the framing
viable, increasing food production, and providing economical access to food to the
larger population through the launch of “Rashtriya Krishi Vikas Yojana (RKVY)” in
2007. The agricultural GDP was considerably stabilized during this phase. At the
end of the 11th 5-year plan, another important program, the “National Food Security
Mission (NFSM),” was launched to promote the production of rice, wheat, and
pulses in the selected districts of India, which has achieved the targeted production
of each crop [25, 29]. To provide economic access to food, another vital program, the
“National Food Security Act (NFSA)” was launched. Several studies have stated that
after the recent initiatives in the fifth phase, a significant improvement has been
recorded in the agricultural production and country’s agricultural GDP [25, 30].

3 Land Degradation Status in India

As a developing country, India has experienced substantial monetary loss in the


agricultural sector due to intensive land degradation [31]. In 2009, India lost about
5.35 billion USD due to land-use change [9]. As per UNEP’s estimate in 1980,
Agricultural Land Degradation in India 223

Fig. 1 The status of land degradation projected by various organizations of India between 1976 and
2004 [34–40]

India’s agricultural land degradation accounted for 4% of its total agriculture GDP
[32]. Recent studies reported that India experienced a loss of Rs. 3,177 billion
(2.54% of India’s GDP) from land degradation and land-use change in
2014–2015, in which agricultural loss accounted for 23% of the total cost
[33]. Among various soil degradation processes, the economic loss due to soil
erosion by water shares a significant proportion (60%) in the total agricultural
loss [33].
The extent of land degradation in India has been previously examined by various
organizations [34–40] according to the land-use statistical report obtained from
secondary estimates (Fig. 1) [41]. In 1976, the National Commission on Agriculture
estimated that 148.09 Mha of the total geographical area was degraded [34]. In 1978,
1984, 1985, and 1994, the Ministry of Agriculture and Society for Promotion of
Wastelands Development (SPWD) projected that around 175, 129.58, 173.64, and
107.43 Mha of land were degraded. These organizations estimated the extent of land
degradation in early times only based on the analysis of secondary data, and no
systematic way of survey and delineation was done. However, accurate information
on the spatial extent of degraded lands plays a vital role in better management and
may mitigate the negative impacts of land degradation on the agro-economy
[42]. The recent advancements in geospatial techniques have allowed policymakers
to examine and monitor the spatio-temporal distribution and changes in the land use
and land cover accurately [43]. The national-level assessment and mapping of the
extent of land degradation were initiated in the 1980s on a 1:1 million scale in the
224 S. Periasamy and R. S. Shanmugam

early phase of exploring remote sensing data applications [40, 44]. According to the
NBSS & LUP report [40], in many research studies, it was stated that nearly 30%
(147 Mha) of agricultural soil has been subjected to degradation from 2004 to date.
With the pioneering research on satellite data acquisition and image processing
techniques in the early 2000s, India has achieved its excellence in mapping the
distribution of land degradation for both regional and local scale at a much finer
resolution [45].
A scientific report on the extent of land degradation in India was prepared and
published by Space Application Centre (SAC) in association with various state
remote sensing centers between 2003 and 2005 and between 2011 and 2013 based
on temporal analysis of Indian remote sensing satellite data by adopting standard
mapping protocol. According to the report [46], published in 2016, out of 328.7 Mha
of India’s total geographical area, nearly 96.4 Mha (28.76%) of land was declared as
degraded land (Fig. 1) on a 1:500,000 scale. The leading cause of the agricultural
land degradation is soil erosion by water (10.98%), followed by vegetation degra-
dation (8.91%) and wind erosion (5.55%). Other studies have also confirmed that out
of 142 Mha of net sown area, 64 Mha of land is affected by various degrees of land
degradation processes [47].
The trend for land degradation’s extent shows that the distribution of degraded
land has declined from 1994 to 2013 (57–29%). This appears to be the result of
increased public investments to address the land degradation problems after 1996
[9]. Due to the high priority given to the restoration of degraded lands with the aim of
achieving the Bonn challenge, India has restored 9.8 Mha of degraded lands between
2011 and 2018 as reported by the International Union for Conservation of Nature
and Natural Resources (IUCN) [48]. The dynamic spatio-temporal trend of the
extent of the land degradation was the evidence for the parallel conversion of
degraded lands into the arable lands and arable lands into wastelands. Despite
intensive land degradation processes and subsequent agro-economic failure, a sig-
nificant increase in the cropland category (5 Mha) was recorded between 2001 and
2009 [9]. The change analysis performed between 2003–2005 and 2011–2013
revealed that 1.95 Mha of land had been reclaimed and 0.4 Mha of land has been
shown improvement in few states of India [46]. Despite this, the change matrix also
revealed a slight increase (0.57%) in a total area undergoing degradation. While few
states of India, namely Delhi, Tripura, Nagaland, Himachal Pradesh, and Mizoram,
have considerably lost the cropped area, states like Odisha, Rajasthan, Telangana,
and Uttar Pradesh have shown improvement. The reduction in the percentage of
degraded land in few places has demonstrated the success rate of the reclamation
measures in those areas.

4 Land Degradation Status in the Climatic Zones of India

India is well known for various climatic regions, namely arid, semi-arid, sub-humid,
and dryland (Fig. 2).
Agricultural Land Degradation in India 225

Legend
Arid
Semi-Arid
Sub-Humid
Other

Fig. 2 The three major climatic regions of India are shown according to Köppen climate classi-
fication system. The classification was done by identifying the relationship between the climate
(mean monthly and annual temperature and precipitation) and the distribution of vegetation
(Original source: National Bureau of Soil Survey and Land Use Planning, Bangalore, Map credit:
Desertification and Land Degradation Atlas of India by SAC, ISRO, 2016)

India’s arid part is dominantly affected by wind erosion process, and the degra-
dation in the semi-arid and sub-humid areas is chiefly triggered by water erosion
followed by vegetation degradation [46]. The arid regions comprise two zones,
namely hot arid and cold arid, which cover 12 and 2% of India’s total geographical
area, respectively. The hot arid zones are majorly spread over the western part of
India, including the parts of Rajasthan (61%), Gujarat (20%), and Punjab and
Haryana (9%) [49], while cold arid areas are distributed in the northern part of the
country that mainly covers Ladakh district (72%) of Jammu and Kashmir State
[50]. Wind erosion is the major degradation process that affects nearly 17.63 Mha of
productive land in the country’s hot arid areas [44, 51]. On the other hand, in the cold
arid region (covering 2% of the country’s total geographical area), around 3.78 Mha
of land is severely affected by frost shattering and mass movement
226 S. Periasamy and R. S. Shanmugam

Fig. 3 The contribution of various land degradation processes in the arid, semi-arid, and sub-humid
areas of India is projected as per the statistics published in Desertification and Land degradation
Atlas of India by SAC, ISRO in 2016

[52]. Approximately 10% of the hot arid areas lies between a north-south zone of
Deccan Plateau comprising the states Andhra Pradesh, Karnataka, and Maharashtra,
in which about 8.41 Mha of land was affected by processes like water erosion
(3.30 Mha), vegetal degradation (2.86 Mha), and salinity (2.52 Mha) [46]. Figure 3
indicates the summary of the land degradation processes in India’s arid, semi-arid,
and sub-humid areas.
The semi-arid area covers nearly 37% of the country’s total area that extends
majorly in the states of Maharashtra (19%), Karnataka (15%), Andhra Pradesh
(15%), Rajasthan (13%), Tamil Nadu (10%), Gujarat (9%), Uttar Pradesh (7%),
Madhya Pradesh (6%), and Punjab & Haryana (6%) [53]. The agricultural activities
are extensive in the semi-arid regions subjected to soil type and water availability. In
these areas of India, the dominant degradation process that threatens agricultural
production is water erosion [46]. About 17.51 and 13.48 Mha of land in the semi-
arid areas are severely affected by water erosion and vegetal degradation, respec-
tively (Fig. 3). The arid and semi-arid regions are subjected to low rainfall reception
and poor vegetative cover, which results in more soil loss than the humid regions.
The north-eastern states cover the maximum proportion of the sub-humid areas of
India, which are affected by water erosion followed by vegetal degradation. How-
ever, compared to the semi-arid regions, the area under degradation by water erosion
(8.97 Mha) and vegetation loss (6.65 Mha) is considerably less (50%) in the
sub-humid areas irrespective of the good annual rainfall reception. This is due to
the presence of extensive vegetation cover that arrests the soil from being eroded
from its primary place and adaptation of various mitigation measures. In these areas,
soil erosion by water is often facilitated by virtue of topography and high rainfall.
Agricultural Land Degradation in India 227

5 Land Degradation Processes in India

5.1 Soil Erosion and Overgrazing

Soil erosion is a natural destruction process of removal of topsoil with high organic
matter and nutrient content driven by the chief natural agents like wind and water,
which is often favored by geomorphology of the site [54], removal of bio-shield
from the field by overgrazing [55], tillage practice [56, 57], and deforestation [58]
(Fig. 4). Removal of fertile topsoil poses a significant threat to crop production.
During the erosion process, soil particles are detached and removed from their
primary location. The water erosion process not only affects crop production but
also tends to pollute the adjacent low-lying areas. Among the two natural agents, i.e.,
wind and water, soil erosion is majorly driven by water in India.
Livestock production is the primary source of income for the people in arid and
semi-arid regions in India. Hence, the increased cattle population in these regions
causes severe overgrazing, resulting in severe soil erosion, which is 3–18 times
greater than the average occurrence at the macroscale [55]. Among all the crucial
land degradation processes of India, overgrazing has a strong connection with the
soil erosion process in the arid regions [55] causing land degradation in eight Indian
states [5]. The soil erosion cycle is explained in the Indian context in Fig. 4. This
process exhibits remarkable spatio-temporal variation due to climate, topography,
land-use/land-cover changes, and intensive agricultural practices. Hence, under-
standing the factors responsible for accelerating this process in different climatic
conditions is of utmost importance before it jeopardizes agricultural productivity
considerably.
In India, about 13.3% of the land has the minimum rate of soil loss tolerance limit
(2.5 t/ha/year) and is prioritized as a high-risk category under soil erosion by water
[59]. It has been reported in several studies that about 37 Mha of the country’s land is
experiencing severe soil loss (40–80 t/ha/year) due to water erosion, and the highest
has been recorded in Madhya Pradesh, where around 26,420 t/ha of soil is lost
annually [40]. Recent studies reported that about 13.4 Mt of crop production,
including cereal, oilseeds, and the pulse, is lost due to soil erosion [31].
The agricultural land degradation in India is majorly witnessed by soil erosion.
About 61% of the total eroded soil is transferred from its primary place, 29% is
deposited in the sea, and the residual 10% is dropped in reservoirs [5]. Nearly
54 Mha (16.5% of the total geographical area) of land is severely affected by soil
erosion, out of which 37 Mha is affected by water, and 18 Mha is affected by wind
[46]. The area affected by water erosion has increased about 0.5 Mha since
2005–2013 [46], with this process being responsible for the severe vegetal degra-
dation of 25 states of India, contributing 37.4% of the total degradation (96.4 Mha).
The Indian states Maharashtra (8 Mha), Karnataka (5 Mha), Orissa (4.4 Mha),
Jharkhand (4 Mha), Gujarat (3.9 Mha), Telangana (2.9 Mha), Rajasthan
(2.1 Mha), West Bengal (1.3 Mha), and Madhya Pradesh (1.1 Mha) are predomi-
nately affected by water erosion (Fig. 5). Further, a few eastern states of India are
228

TemperatureVegetation type

Wind

Precipitation Soil Erosion

by
Human activity
wind and water
Topography

Overgrazing
Soil type

Tillage
Unsustainable
agricultural practices Top soil loss

Precipitation
Eutrophication
Human pressure
on agricultural land

Runoff Vegetation loss

Desertification
over a time

Accelerated soil erosion


Land degradation

Fig. 4 Soil erosion cycle in the Indian context. Soil erosion processes are activated by natural agents like water and wind with the support of climate,
S. Periasamy and R. S. Shanmugam

topography, soil type, and human practices. The removal of topsoil affects vegetation production, and fertilizers from the soil to the water bodies lead to
eutrophication. Vegetation loss and eutrophication increase runoff that accelerates soil erosion and results in land degradation. Further, human activities act as a
threat multiplier for the already degrading soil, thus makes it even more vulnerable to the erosion process
Agricultural Land Degradation in India 229

Fig. 5 The percentage contribution of various degradation processes is projected state-wise as per
the statistics published in Desertification and Land degradation Atlas of India by SAC, ISRO
in 2016

also actively affected by water erosion, particularly in the state Tripura and Megha-
laya where about 42 and 10.7% of the total degraded land is affected by water
erosion because of reception of quantum of rainfall and sloped cultivation practice
[9, 46].
Wind erosion is the most prominent process in India’s hot arid regions, respon-
sible for the loss of total organic carbon in the soil. Nearly 660 Mt of soil is being lost
every year in India due to wind erosion [60]. Wind erosion actively removes the
topsoil (where the nutrients and bacterial activity is higher), majorly from the desert
areas of India. The process affects the soil’s stability, thereby reducing the water
holding capacity of the soil and nutrient availability to the crops [61]. The process is
proactive in the north-western part of the country, covered by desert areas. Out of
18 Mha of total affected land by wind erosion, nearly 83% (15 Mha) of land is shared
by Rajasthan and the adjoining areas of Jammu and Kashmir (9%) and Gujarat (6%).
According to the scientific report of SAC, the land area affected by wind erosion has
decreased by 0.1 Mha from 2005 to 2013 due to the implementation of various wind
control measures by the Indian government.
230 S. Periasamy and R. S. Shanmugam

5.2 Salinity/Alkalinity

Next to soil erosion by water and wind, a relevant process accountable for severe
land degradation in the nine states of India is salinization/alkalinity (Fig. 6) [46]. The
ever-expanding extent of land degradation due to soil salinity has become a severe
threat to agricultural production and economic growth. Soil salinity is a typical land
degradation process in arid and semi-arid environments where evaporation is higher
than precipitation [62]. In India, around 3.8% (3.7 Mha) of the total degraded land
(96.4 Mha) is affected by salinity [46]. Of 3.7 Mha of total degraded land, 2.65 Mha
(72%) is distributed across Gujarat state alone. Reports state that the areal extent of
the salt-affected regions is kept increasing by 10% annually in India [63], and if it
continues, it is estimated to be increased to 16.2 Mha by 2050 [64]. Further, nearly
5–11 Mha out of 141 Mha of irrigated land was estimated to be converted into barren
land due to primary and secondary salinization by 2050 [65–67]. Primary salinity
occurs in agricultural soil naturally supported by the upward flux of groundwater
with capillary flow from the shallow aquifer in the pre-monsoon period [68, 69],
whereas secondary salinity is a result of human activities for intensive agricultural
practices [70].
Based on the SAC report (2016) on the overall status of the saline-affected lands
of India, the area affected by salinity decreased by 0.3 Mha between 2005 and 2013.
Being a Peninsula country surrounded by three oceans, i.e., the Arabian Sea on the
west, the Indian Ocean on the south, and the Bay of Bengal on the east, the
agricultural lands of coastal states of India like Gujarat (2.65 Mha), Andhra Pradesh
(0.12 Mha), Karnataka (0.09 Mha), Maharashtra (0.03 Mha), and Tamil Nadu
(0.01 Mha) are undergoing degradation by coastal induced salinity problems due
to the proximity to the sea [46, 71]. The coastal salinity may degrade the soil by a
wide variety of processes. The groundwater at shallower depth quickly gets con-
taminated in the coastal area by seawater intrusion due to the over-extraction of
groundwater for irrigation (Fig. 6). The salinity also reaches the soil root zone by
using contaminated groundwater for irrigation purposes, the salts in the groundwater
take upward migration through capillary action and reach the crop’s root zone
(Fig. 7). The speed of such salt migration is subjected to the soil type and temper-
ature. In Rajasthan, soil salinity has been associated with the use of contaminated
groundwater for irrigation [72] and the absence of natural drainage [73]. The
capillary rise is considerably lower in sandy soil than the clay soil in a deeper
depth. Since the soil type in this state is predominately sandy in texture and
groundwater’s average depth is 23.58 m below ground level [74], the upward
migration of salt solutions from shallow groundwater to the soil root zone is
considerably controlled.
Soil salinity in the irrigated land can also be caused by excessive application of
fertilizers and other chemicals (Fig. 6). This land degradation process may produce
an irreversible effect on the soil if it continues over time. India has experienced a
monetary loss of Rs. 230.20 billion annually by losing 16.84 MT of agricultural
production due to soil salinity [67]. Many research institutes in India are engaged in
Ocean

Precipitation
Ground water intrusion

Irrigation Practice

Waterlogging

Soil salinity
Agricultural Land Degradation in India

Ground water recharge

Crop failure

Reduction in osmotic
Chemical Fertilizers
potential of plants

Rock-water Capillary rise


interaction

Temperature

Fig. 6 Soil salinity cycle in Indian context. Being a country surrounded by three oceans, with diversified climate and geologically saline-rich minerals, Indian
soil is affected by both seawater and groundwater salinity. Soils near the coastal region are affected by saltwater intrusion due to over-extraction of groundwater,
and soils in the geologically active area are affected by groundwater salinity through capillary action. The application of chemical fertilizers and irrigation
practices also add salinity to the soil, resulting in waterlogging and reduction in the osmotic potential of plants
231
232 S. Periasamy and R. S. Shanmugam

Fig. 7 Unproductive land due to higher accumulation of salinity at the root zone, location: Vellore
district of Tamil Nadu, India, date of observation: 20th July 2019

research related to salt-affected soils and investigating the appropriate reclamation


measures [67]. To ensure food security and environmental sustainability, the Gov-
ernment of India has fixed a goal to restore around 26 Mha of degraded land by 2030
[63]. So there is a need to understand the dynamic behavior of the salinity process
and to identify the appropriate land management technique to address food security
for the expanding population and monetary loss.

5.3 Waterlogging

The country’s annual agricultural growth is projected only about 2.6% over the last
25 years [75]. To reduce the problems in the farming sector due to frequent monsoon
failure, the Government of India has developed a canal irrigation plan for about
57 Mha of arable land [76]. As a result, irrigated lands were mainly affected by the
twin effects of salinity and waterlogging caused by excessive irrigation of the canal
water in the conception of obtaining more yield [77]. Because the irrigation water
also carried salt concentration (0.3 dS/m) [77], it was reported that around 5.5 Mha
of the saline-affected land is waterlogged [75]. In Fig. 8, the agriculture failure due to
irrigation-induced waterlogging in the saline-affected area is shown.
Waterlogging is a dynamic process in which the irrigated water saturates the soil
surface and tends to be stagnated as a layer for a longer time in the absence of a
natural or artificial drainage system. The waterlogging condition in the saline-
Agricultural Land Degradation in India 233

Fig. 8 Agriculture failure due to waterlogging condition. Location: Perambalur district of Tamil
Nadu, India, date of observation: 25th November 2018

affected patches is mainly due to the virtual absence of drainage, low precipitation,
and high temperature [78]. Waterlogging in the agricultural areas is also a result of
water table rise by capillary action and unavailability of natural and artificial
drainage system in the soil. The continued existence of crops is severely constrained
by waterlogging as it restricts the air circulation in the plants’ root zone. The
restricted air movement in the saturated zone results in a reduced level of O2 and
an increased rate of CO2 [79]. The combined effect of soil salinity and waterlogging
is considered a significant problem in India’s irrigated lands [80]. In India, the flood-
prone regions [81], areas with shallow groundwater and high rainfall [82], coastal
areas with saltwater intrusion [83], and irrigated command areas [84] are under high
risk of waterlogging hazard.
In 1976, the National Commission on Agriculture identified nearly 6 Mha of land
was waterlogged, out of which 3.4 Mha of land was affected by surface water
stagnation, and 2.6 Mha of land was affected by water table rise [85]. According
to the Ministry of Agriculture report, during 1984–1985, an area of 8.53 Mha of
land, including irrigated and unirrigated lands, was estimated to have been affected
by the waterlogging problem. According to the report published by NRSC in 2005,
the surface stagnation of water affects about 1.66 Mha of wastelands while
sub-surface waterlogging affects 4.75 Mha of India’s agricultural land [44]. A recent
macro-scale study conducted in 2013 showed that 0.7 Mha of India’s agricultural
land was subjected to waterlogging conditions [46]. In the eastern states of India, the
problem is significant, especially in Assam, where nearly 186,667 ha of land is
affected by waterlogging due to excessive irrigation, while in the southern part of
India like Andhra Pradesh, the twin effects of salinity (117,952 ha) and waterlogging
(132,334 ha) are prominent. An overall increase of about 9% in waterlogging areas
was observed between 2005 and 2013.
234 S. Periasamy and R. S. Shanmugam

5.4 Agrochemicals

With a total of 26.5 Mt per year, India ranks second in the world in the consumption
of fertilizers [86]. The application of fertilizers requires a scientific approach based
on the soil testing results. But, at present, Indian agriculture has a negative balance of
necessary plant nutrients because of the use of fertilizers without knowing the actual
requirement of the soil [87]. The intensive use of agrochemicals on agricultural land
was encouraged in countries like India during the Green Revolution to achieve self-
sufficiency in food production and to support the growing population. In the initial
period of the Green Revolution, a positive effect of fertilizers was reported in terms
of agricultural production. However, due to the prolonged use of chemical pesticides
and fertilizers, there were negative consequences on the soil, like declining rate of
microorganisms and significant change in the physico-chemical characteristics of the
soil [88]. The excessive pressure on agricultural land in the conception of crop
diversification, increasing the crop yield, short-time benefit, land shortage, and
intensive farming practices resulted in inappropriate land-use policies and manage-
ment [89]. Indian farmers are started using agrochemicals to increase the productiv-
ity, obtain the expected yield from the wide variety of seeds, and overcome the
deficiency in micro- and macro-nutrients in the soil. The unbalanced use of such
agrochemicals beyond the limit causes an adverse effect on soil fertility character-
istics, thereby impacting crop production. The application of fertilizers beyond the
assimilative capacity of soil adversely affects the nearby environment and enhances
the soil’s erodibility rate.
The fertility status of the soil distributed in the Indo-Gangetic plain shows a
declining trend due to the application of agrochemicals [87]. The northern states of
India are likely to use excessive fertilizers to produce cereal crops and eventually
contaminate soil and the nearby water sources [9]. During 1990–1991, 2000–2001,
2009–2010, and 2014–15, the rate of imbalanced fertilizer use was widened from
6.2:4:1 (N: P: K) to 31.4:8.0:1 (N: P: K), well above the desirable limit 4:2:1,
especially in the agriculturally important states Punjab and Haryana [5, 87]. These
states are often recognized for their significant contribution to rice and wheat
production, which have virtually extracted essential nutrients from the soil. About
292 out of the 525 districts in the country accounted for 85% of the total fertilizer
consumption. The continuous application of agrochemicals has reduced the required
proportion of sulfur, zinc, and boron in many soils of the country. The minerals’
unequal balance has affected the soil response to fertilizers and the crop
productivity [87].
Furthermore, agrochemicals like fertilizers, pesticides, and disposal of effluents
from the industries result in significant soil damage and are often responsible for land
degradation. The disposal of effluents and application of fertilizers often resulted in
the accumulation of heavy metals like arsenic (As) and fluoride (F) in the soil and
water that causes undesirable effects on the crop production and human food cycle
(Fig. 9). Despite overconsumption of the major essential nutrients, comprehensive
deficiency of Zn, followed by S, Fe, Cu, Mn, and B, was reported throughout the
country [5].
Agricultural Land Degradation in India 235

Fig. 9 The use of fertilizers and disposal of industrial effluent into the arable land. Location:
Vellore district of Tamil Nadu, India, date of observation: 15th January 2020

5.5 Slash-and-Burn Agriculture

Another crucial farming method that contributes to land degradation in India is the
slash-and-burn agricultural practice. The slush-and-burning method, also known as
Jhum cultivation and stubble burning in India, involves cutting and burning the
plants as shifting cultivation, deforestation, and post-harvest practice, which
adversely affects the soil organic content to a more significant extent by killing the
microorganisms that are essential for soil to balance the fertility characteristics
(Fig. 10). It was estimated that the burning of one ton of rice crop residue resulted
in the loss of essential nutrients like nitrogen (5.5 kg), phosphorus (2.3 kg), potas-
sium (25 kg), and sulfur (1.2 kg) [90]. The demand for crop residue for cattle feed
and industrial activities is ever-increasing in India because of the in-situ crop burning
practices. The rice-wheat system cultivation accounted for nearly 70% of India’s
total crop residues production [91]. According to the statistics published by the
Ministry of Statistics and Programme during 2013–2014, nearly 516.3 Mt of crop
residue generated annually in India, out of which 110 Mt is shared by wheat, 122 Mt
by rice, 71 Mt by maize, 26 Mt by millets, 141 Mt by sugarcane, 8 Mt by fiber crops,
and 28 Mt by pulses. The highest crop residue production was recorded in the state
Uttar Pradesh (60 Mt), followed by Punjab (51 Mt) and Maharashtra (46 Mt). Of the
total crop residue production, nearly 21% (109.2 Mt) and 10% (52.7 Mt) were
contributed by Uttar Pradesh and Maharashtra states, respectively [92]. In both of
these states, the major share of crop residue was contributed by sugarcane. A paddy
waste contributed 40% of the total residue burnt in India since 2008–2009 [93]. Of
various crop residues, wheat, rice, and sugarcane crop residues were involved in crop
burning in the north and southern part of India [92, 94].
Out of the 501.76 Mt of stubble generated annually [95], 92.81 Mt of the stubble
is being burned on-site in which the highest being recorded in the state Uttar Pradesh
(21.92 Mt) followed by Punjab (19.62 Mt), Haryana (9.06 Mt), and Maharashtra
(7.41 Mt) [95, 96]. The southern states like Karnataka (6%) and Tamil Nadu (4.4%)
236 S. Periasamy and R. S. Shanmugam

Fig. 10 The agricultural land patch is shown with sugarcane crop residues burnt after harvest.
Location: Namakkal district of Tamil Nadu, India, date of observation: 13th October 2018

also contribute to burn agriculture. According to the Ministry of Statistics and


Programme, in India, the trend of crop residue burning has been increasing by
2.53% annually. It is also reported that the amount of residue burned in India is
approximately equal to the total amount of crop residue production by other coun-
tries like Bangladesh, Indonesia, and Myanmar [97]. Although there are multiple
productive alternatives to the crop burning, nearly 3.7 Mha of agricultural land in
India has been found to hold low soil organic matter contents due to crop residue
burning [40]. This is because that the farmers consider that burning crop residues is
the only option to remove the residue from the farmland and prepare the land for the
next sowing due to the lack of awareness on the profitable alternatives [98]. The crop
residues greatly help reduce the impact of other soil degradation processes like
salinity and erosion and increase the soil fertility in all its dimensions if sustainably
managed.

5.6 Microplastics

Microplastics (synthetic polymer-based particles with less than 5 mm size) are one of
the serious and emerging pollutants of India’s agricultural land as it has been
increased exponentially in the past years [99]. According to the Central Pollution
Control Board (2015) statistics, nearly 62 Mt of solid waste was produced in India,
out of which about 36 Mt of waste was openly dumped without treatment causing
Agricultural Land Degradation in India 237

water and soil pollution [14]. Microplastics endanger the agricultural production in
India as it induces significant changes in the biophysical and chemical properties of
the soil. Further, the earthworm’s life is under severe threat due to the microplastics
that significantly degrade the fertility of the agricultural soil. In India, it is reported
that the microplastics reach the soil in the form of nutrient prills for the controlled
release of nutrients to the soil, capsules for crop production, water-soluble polymers
for soil remediation, water absorbents, and seed coating [100]. The accumulated
plastic debris in the soil undergoes macro- and micro-scale breakdown by photo- and
bio-degradation process. While microplastics in the ocean and freshwater ecosystem
have been widely studied, their consequences on agricultural soil productivity are
overlooked [101–103]. Recent studies have also proved that agricultural soil has a
higher potential to store microplastics (>40,000 particles kg-1) than the ocean
basins [104]. Microplastics in agricultural soils are primarily “secondary
microplastics” (with the fibers as the predominant element (92%) followed by
fragments (4%)) derived from the disintegration of larger plastic particles in the
environment. Microplastics commonly reach the soil by anthropogenic sources
represented by agricultural activities like plastic mulching, irrigation using plastic-
polluted water, and soil amendment practice with plastics [105].

5.7 Fluoride

Among all other contaminants, the important geogenic pollutant which threatens
agricultural sustainability and is found to be responsible for the severe soil degra-
dation in 19 states of India is fluoride (F) [106, 107]. The distribution of F roughly
shares more than half of the country’s total geographical area. F is the most
widespread pollutant in India, followed by As, as the country shares a significant
percentage (14%) of the world’s total F deposition [108, 109]. In India, there are
three common types of F found in the soil, namely calcium fluoride (CaF), sodium
fluoride (NaF), and hydrogen fluoride (HF) as natural and industrial versions in
which NaF and HF are majorly distributed and found to be very harmful to the
environment and the health of the human community. This pollutant is widely
distributed in various ranges in all the environment components, such as air,
water, soil, plant, and rocks [110]. The intake of F by plants from soil and the
associated health risks are more significant in India than in other countries [111]. The
F contamination in agricultural soil is kept increasing ever since, and however, its
adverse effects on soil’s physical and chemical properties are being given less
attention. While concentrating mainly on the impacts of major soil degradation
processes on biomass reduction, F’s contribution to agricultural soil degradation
and associated human and livestock health risk is often overlooked in India. The
concentration of F is relatively higher in clay soil than sandy soil [112] as the rate of
F release increases in the soil which is capable of exchanging cations [113]. In low
pH ranges, the F present in the clay soil as a mineral due to electrostatic potential. In
contrast, in the higher pH range, it is available in soil solution due to the alkaline
238 S. Periasamy and R. S. Shanmugam

condition that favors the dissolution of F and repulsion between soil particles and F
ions in the negatively charged surfaces [114]. So, in the low pH values, the soil
transfers the contamination directly to the crops, while in the higher pH values, the F
triggers crop-water stress by reducing the plant’s osmotic potential.
The deposition and accumulation of F in the soil’s root zone results from natural
and anthropogenic causes. But in the Indian context, the contaminated groundwater
by country rocks is considered a major source of F supply to the soil. Conversely, it
was also reported that soils contaminated from industrial sources played a significant
role in contaminating groundwater [115] through downward leaching. The transport
and deposition of this contamination in the soil profile depends on the soil’s various
physico-chemical conditions and climatic variables. So, as discussed, the naturally
occurred F shows an immediate effect on the groundwater and contributes to
degrading the soil at a slower rate. Likewise, the F release to the environment by
human activities primarily affects the soil properties and subsequently the ground-
water. Hence F cyclically contaminates the soil and groundwater and becomes a part
of human and livestock’s food chain through bioaccumulation. This section
describes the F contamination cycle between soil and groundwater affected by
different sources (Fig. 11).
Weathering and leaching of F-bearing minerals due to rock–water interaction in
the aquifers is the common cause for the F enrichment in the groundwater [116]. This
groundwater F experiences upward movement often from shallow aquifers through
capillary action and reaches the soil profile, favored by climatic conditions, soil type,
soil depth, and water table depth. This process of upward migration through capillary
rise is often accompanied by groundwater salinity. Similarly, F contamination from
anthropogenic activities like application of F inducing phosphate fertilizers and
pesticides [114], groundwater irrigation [117], releasing industrial F-rich effluents,
and industrial emission [118] may reach the soil profile (1) as a dry deposition and
remain as a semi-permanent reservoir of F, (2) by decomposition of the contami-
nated plant tissues, and (3) by atmospheric precipitation [119]. These processes are
further accompanied by the transport of F in water-soluble fractions through the
saturated zone of soil in recharging groundwater through precipitation [120],
resulting in increasing groundwater F [108].
This ever-ending mobilization cycle of F through the atmosphere, soil, and
groundwater may produce a notable effect on the existing ecosystem. The significant
effects of these processes are:
1. F contamination considerably influences the fertile characteristics of the agricul-
tural soil, thereby decreasing crop productivity.
2. The human and livestock are directly exposed to high risk by inhaling the F
released as a gaseous effluent from industries.
3. The accumulation of F in the plant tissues through atmospheric precipitation
affects the yield potential.
4. Irrigating the soil with contaminated water transfers F to the crops, and this
contamination becomes a part of human and livestock’s food chain through
bioaccumulation.
Agricultural Land Degradation in India

Fig. 11 A simplified description of the fluoride (F) migration cycle and its causes and effects. The F contamination reaches the soil through natural and
anthropogenic channels, including the dissolution of F-rich minerals in the groundwater, groundwater recharge with F-contaminated water, chemical fertilizers,
industrial release (air and water), and irrigation practice. The contaminated soil transfers F to the crops, and this contamination becomes a part of human and
239

livestock’s food chain through bioaccumulation. The other direct sources of F to humans and cattle populations include (1) gaseous effluent from industries,
(2) contaminated drinking water, and (3) atmospheric precipitation. The effect of F includes dental and skeletal fluorosis in humans and livestock
240 S. Periasamy and R. S. Shanmugam

6 State-Wise Land Degradation Status

The total agricultural land degradation status was projected state-wise as per the
record published by SAC in 2016 (Fig. 12). Maps were produced on a 1:500,000
scale using Indian Remote Sensing data products [46]. In the western and central
states of India, including Rajasthan, Gujarat, and Madhya Pradesh, more than 10%
of the total geographical area was in a degraded condition. In contrast, in the eastern
states, the distribution of degraded land is minimal compared to its total area. Wind
erosion is dominant in Rajasthan, and water erosion is active in Gujarat and Madhya
Pradesh. Nearly 8% of the total area in the northern state Jammu and Kashmir is
degraded majorly by frost shattering. Karnataka state (southern part of the country)
is reported to have a high proportion (7%) of degraded land due to water erosion. The
state-wise distribution of degraded land under various categories is listed in Table 1.

7 Mitigation Strategies Adopted in Agro-Climatic Zones of


India

The planning commission of India has divided 15 broad agro-climatic zones,


namely, (1) Western Himalayan Region, (2) Eastern Himalayan Region, (3) Lower
Gangetic Plain Region, (4) Middle Gangetic Plain Region, (5) Upper Gangetic
Plains Region, (6) Trans-Ganga Plains Region, (7) Eastern Plateau and Hills,
(8) Central Plateau and Hills, (9) Western Plateau and Hills, (10) Southern Plateau
and Hills, (11) Eastern Coastal Plains and Hills, (12) Western Coastal Plains and
Ghats, (13) Gujarat Plains and Hills, (14) Western Dry Region, and (15) Island
Region in the seventh 5-year plan (1985–90) as an extension of climatic zones to
formulate integrated development plans for agriculture. In this mid-term appraisal
plan, more emphasis was given on the optimum utilization of the resources. The
agro-climatic zones are broadly delineated, considering the land suitability to agri-
culture based on the various climate and crop parameters and to draw extensive
agricultural plans, and to frame the strategies to mitigate land degradation processes.
To date, there are numerous mitigation measures developed and deployed to address
the land degradation problems in the different agro-climatic zones of India. Since the
region’s susceptibility to soil erosion varies from one area to another, appropriate
site-specific mitigation measures need to be selected to combat the problems effec-
tively. This section gives the reader an idea about the different conservation practices
followed in India’s different soil conservation regions to conserve the agricultural
land from the degradation processes like soil erosion, salinity, waterlogging, agro-
chemical, slash-and-burn agriculture, microplastics, and fluoride.
Agricultural Land Degradation in India 241

Fig. 12 Total area of agricultural land degradation projected state-wise as per the report published
by SAC in 2016

7.1 Himalayan Region

The Himalayan Region of India consists of two broad zones, namely the Eastern and
Western Himalayas region. The Indian states under Eastern Himalayas are
Arunachal Pradesh, Assam, Manipur, Meghalaya, Mizoram, Nagaland, Sikkim,
Tripura, and West Bengal (Hilly region), and Western Himalayas are Jammu and
Kashmir, Himachal Pradesh, and Uttaranchal. Recent studies have reported an
alarming rate of soil erosion (>40 t/ha/year) in nearly 39% area of the total
Himalayan region due to its underlying topography and rainfall [121]. The Himala-
yan regions are also severely affected by chemical degradation processes like soil
acidity. The extent of acid-affected soil in the North-Eastern Himalayan Region
(29%) was found to be higher than in the North-Western Himalayan Region (0.8%)
[5]. The North-Eastern parts of Himalaya are also affected by shifting cultivation
Table 1 State-wise land degradation status of India in 2013
242

Water Wind Water- Frost Mass Human-made Barren/


Vegetation erosion erosion Salinity logging shattering movement (inclusive of rocky Total area
State loss (ha) (ha) (ha) (ha) (ha) (ha) (ha) settlements) (ha) (ha) (ha)
Andhra 1,164,257 789,433 3,986 117,952 132,334 – – 70,274 20,521 2,298,758
Pradesh
Arunachal 120,499 – – – – 20,186 – 13,247 – 153,933
Pradesh
Assam 471,958 31,424 – – 186,667 – – 26,548 – 716,596
Bihar 242,525 321,175 – – 106,628 – – 24,480 – 694,809
Chhattisgarh 1,348,089 783,645 – – – – – 72,197 7,222 2,211,153
Delhi 9,980 – – 347 – – 79,541 – 89,868
Goa 138,172 33,889 – – 9,005 – – 11,907 – 192,973
Gujarat 2,319,826 3,859,497 1,177,105 2,645,405 3,375 – – 217,215 39,218 10,261,641
Haryana 41,411 13,568 151,797 27,841 12,530 – – 91,817 – 338,964
Himachal 1,790,803 268,261 – – – 332,423 – 2,753 – 2,394,240
Pradesh
Jammu and 1,951,000 146,932 1,670,244 – 70,563 2,968,279 927,986 15,924 218,679 7,969,607
Kashmir
Jharkhand 1,379,038 4,036,785 – – – – – 82,903 – 5,498,726
Karnataka 1,712,386 5,043,041 2,159 86,740 – – – 103,285 3,389 6,951,000
Kerala 337,613 – – 11,989 – – 29,984 – 379,587
Madhya 2,523,801 1,125,418 – – 7,788 – – 115,813 31,495 3,804,315
Pradesh
Maharashtra 4,884,005 8,060,753 – 29,089 – – – 345,925 506,163 13,825,935
Manipur 575,603 8,070 – – 5,026 – – 13,260 – 601,959
Meghalaya 435,527 53,149 – – 1,548 – 4,656 – 494,880
Mizoram 167,050 8,119 – – – – – 12,285 – 187,453
Nagaland 778,421 – – – – – 8,257 – 786,678
S. Periasamy and R. S. Shanmugam
Orissa 745,122 4,409,413 – – 36,439 – – 108,012 5,128 5,304,114
Punjab 32,561 14,116 – – – – – 97,976 – 144,653
Rajasthan 2,606,221 2,116,314 15,197,874 363,768 18,421 – – 171,540 1,052,374 21,526,512
Sikkim 74,318 – – – – 3,730 – 700 – 78,749
Tamil Nadu 1,385,478 6,411 30,429 9,878 – – – 111,188 515 1,543,898
Telangana 541,145 2,854,285 – 86,514 – – – 114,933 1,979 3,598,856
Tripura 236,374 186,900 – – – – – 13,854 – 437,128
Uttar 413,476 586,961 – 307,571 33,620 – – 187,369 – 1,528,997
Pradesh
Uttarakhand 606,616 11,943 – – – 13,786 – 15,908 – 648,253
West Bengal 265,277 1,329,539 – – 17,627 – – 121,488 – 1,733,931
Agricultural Land Degradation in India

Total area 29,298,553 36,099,042 18,233,594 3,674,759 653,908 3,338,404 927,986 2,285,239 1,886,682 96,398,166
(ha)
Source: Desertification and Land Degradation Atlas of India by SAC, ISRO, 2016
243
244 S. Periasamy and R. S. Shanmugam

activity [122]. Thus, there are three physical and chemical stresses in the Himalayan
region, namely water erosion, soil acidity, and shifting cultivation practices that are
actively involved in degrading the agricultural soil.
More than 72% of the nutrient loss in the agricultural areas of the Himalayan
region is attributed to soil erosion. The extensive agricultural practice and Jhum
cultivation in the steep slopes has led to accelerated soil erosion [123] that increases
the need for proper conservation measures. According to the rainfall distribution
(Northern Himalayan ~500–2,000 mm, and North-Eastern Himalayan
~1,500–2,500 mm) and underlying steep topography, proper mitigation measures
have been identified and implemented including (1) creating contour tillage along
the contours to reduce the velocity of the water and increase surface drainage
(2) creating contour bunding from either vegetative cover or mechanical measures
across the slope, (3) construction of check dams (as a soil and water conservative
measure from integrated watershed approach) to increase the water infiltration and
reduce subsequent soil erosion, (4) encouraging subsoiling and broad-bed furrows
for the vertisols, (5) promoting terrace farming, and (6) adopting strip crop farming
system in which the usual crops are grown along with soil conserving crops on the
same strip perpendicular to the direction of the wind and water [122, 124]. Soil
acidity is often controlled by adding an appropriate dose of lime [125, 126] to the soil
during the time of crop sowing.

7.2 Indo-Gangetic Plains (IGP)

Indo-Gangetic Plains (IGP) comprise four agro-climatic zones, including Lower


Gangetic Plain Region, Middle Gangetic Plain Region, Upper Gangetic Plains
Region, Trans-Ganga Plains Region that covers the states Punjab, Haryana, Uttar
Pradesh, Bihar, and some parts of West Bengal. A higher proportion of productive
soils are distributed in the IGP, and the annual rainfall distribution ranges from
700 to 1,000 mm. The IGP plain is generally referred to as “water-surplus plain” or
“large flood plain” of river systems, namely Indus, Ganga, Yamuna, Ghaghara,
Gandak, and Kosi that are emerging from Himalayas. Canal irrigation was intro-
duced at the end of the nineteenth century in this plain to control the problem of
acidity and increase crop yield. Excessive irrigation has significantly raised the
groundwater water resulted in waterlogging. The rice-wheat farming system in
IGP states is considered the major source of food supply after the successful
implementation of the Green Revolution. However, despite the paramount impor-
tance in India’s total rice production, the deceleration trend in rice and wheat
production has been observed in the recent era [127]. The intensively cultivated
areas under IGP are notably affected by sheet erosion and nutrient imbalance
[122, 128]. The rice-wheat system cultivation in these areas has mined the major
nutrients leading to the depletion of soil organic carbon [129] and eventually
accelerated soil erosion by water over time [124].
Agricultural Land Degradation in India 245

Conservation agriculture is a site-specific mitigation measure commonly adopted


in IGP to minimize soil loss and maximize crop yield, especially for the agricultural
areas undergoing rice-wheat rotation. Conservation agriculture is a farming system
that controls soil loss by adopting three crucial principles such as (1) minimum
mechanical soil disturbance with zero tillage, (2) makeup the permanent soil cover
using crop residue and live mulches to reduce the direct impact on the soil structure,
and (3) encourage crop rotation and intercropping system. There are reports of
higher control on soil loss and productivity in the maize-wheat farming system
with a 2% slope in the areas of Indian Himalayas due to the practice of conservation
agriculture [130]. Recent research reports indicated that the combined operations
like zero tillage and crop cover have shown higher potential in soil conservation and
increasing crop yield in tropical mountainous areas, especially in the eastern part of
IGP (400–1,000 kg/ha) [128]. The implementation of conservation agriculture in one
of the IGP states, Uttar Pradesh, has led to the decrease in degraded land by water
erosion (24,028 ha) from 2005 to 2013. However, the other states of IGP, including
Punjab, Haryana, Bihar, and West Bengal, showed an increasing trend in water
erosion which needs special attention from the governmental bodies. Combining the
advantageous elements of conservation agriculture and integrated watershed
approach has a higher potential to reverse the soil loss [5]. Moreover, the use of
groundwater for irrigation is recommended rather than canal irrigation to lower the
water table level and avoid waterlogging problems in the IGP plains.

7.3 Dry and Arid Regions

Wind erosion is an active degradation process in India’s dry and arid areas, which
has often resulted in multidimensional problems [131, 132]. With an annual average
rainfall of 150–500 mm, soils in the arid regions are often found with structure-less
topsoil with sparse vegetation under dry conditions, making the area more vulner-
able to wind erosion. Wind erosion can be controlled either by reducing the wind’s
speed or by decreasing the susceptibility of the soil to erosion using natural or
artificial barriers against the wind direction. The measures include (1) sand-dune
stabilization by checkerboard and planting of ecologically viable plants based on the
moisture availability, temperature, wind velocity, and salinity conditions, (2) increas-
ing surface cover in the sandy plains either by grass crops or by cover crops, (3) strict
measures to have a controlled grazing practice in the rangelands to retain the
sufficient wind-protective grass cover, (4) use of plant residues to control the effect
of wind erosion, (5) increasing the soil cohesiveness, (6) reducing the momentum of
wind by decreasing the track length of the agricultural field, (7) providing wind-
breaks, (8) establishing bioshields from the trees, particularly Prosopis juliflora,
Cassia siamea, and Acacia tortilis, to check the wind speed and damage, (9) crop
management like planting taller crops (Pennisetum glaucum, Sesamum indicum, and
Ricinus communis) in between the shorter crops (Vigna radiata, Cyamopsis
tetragonoloba, and Arachis hypogaea) to protect them from wind action, and
246 S. Periasamy and R. S. Shanmugam

(10) tillage operation to bring the sub-surface clay soil to the surface layers
[128, 133–135]. Nearly 0.1 Mha of agricultural land was reclaimed from wind
erosion from 2005 to 2013 due to the implementation of extensive conservation
measures.

7.4 Coastal Regions

The coastal agro-ecosystem of India, known for its diverse climatic and soil condi-
tions, shares around 10.8 Mha of the total geographical area. The agricultural
production in the coastal area primarily comes from mono-crops as it highly depends
on rainwater which occurs only in monsoon months. The agricultural regions spread
over both east and west coast region are highly affected by acidification, salinization,
and waterlogging problems [136]. The rainwater harvesting in the pond, also known
as the “land shaping technique,” has proven to increase the productivity of the salt-
affected agricultural lands of coastal India [137]. This technique aimed to address
some of the crucial issues of coastal agriculture, including lack of irrigation
resources, drainage congestion, brackish groundwater at a shallower depth, and
mono-cropping cultivation. The technique greatly helped the farmers in the
Sundarbans (a mangrove area in the delta region at the convergence point of Ganges,
Brahmaputra, and Meghna Rivers in the Bay of Bengal) to combat the prolonged
salinity and waterlogging problems and increase farm productivity. Crop diversifi-
cation was successfully carried out in nearly 370 ha of land in Sundarbans and the
Andaman and Nicobar Islands, adopting land shaping technique that includes farm-
ponds, deep furrow and high ridges, paddy-cum-fish, broad-bed and furrow, the
three-tier system, the paired bed system, and the drainage improvement network
under the GEF funded National Agricultural Innovation Project [138, 139].

7.5 Red and Black Soil

Red and Black soils are widely distributed (~150 Mha) in the central part of and
southern peninsular India, chiefly affected by sheet erosion, waterlogging, salinity,
fluoride contamination, and microplastics. In the sloppy terrain, soil conservation
and drainage were achieved with implementation of the “broad-bed and furrow”
(BBF) system. As a consequence, a considerable percentage of yield has increased,
mainly from legume crops, in the central parts of India [140]. Similarly, the contour
bunding method was deployed in the areas receiving less than 700 mm annual
rainfall and with a slope of less than 6% to decrease the runoff and increase water
retention within the agricultural field [128]. The contour farming practice has
significantly controlled the soil loss (from 18.92 to 0.3 t/ha) and has been reported
to increase 22.3–65.5% of the crop yield for both red and black soil [141].
Agricultural Land Degradation in India 247

The low infiltration problem in central India is being mitigated through the
subsoiling technique. Mitigation measures like providing drainage facilities in
lowlands, recharging the aquifers, and water harvesting in the natural and artificially
created ponds have controlled the rate of water erosion and decreased salinity
occurrence in southern Peninsular India. In a local-scale view, the problem of
salinity can be effectively managed by processes including leaching, irrigation
(with good quality groundwater), and growing salt-tolerant plants. Furthermore,
the supply of NPK mineral fertilizers to the fields, along with organic manure,
increases crop yield and improves soil organic contents [5, 142]. Other than these
methods, few chemical alternatives have been actively used by the farmers to reclaim
the non-saline and non-acidic soils.
In many countries like India, the significant contribution of increased F contam-
ination to soil reclamation failures is often ignored or given less attention. The crop
growth and yield depend on the rate of F deposition in the root zone. To improve
crop growth and prevent the crop from being contaminated with F, it becomes
essential to remove such contamination from the root zone. Implementing the
inappropriate reclamation measures without due consideration of the dynamic
upheaval behavior of contamination and leaching requirement of the soil often
results in an agro-economic loss. Soil refilling is a physical method for soil remedi-
ation in which the F-contaminated soil is removed at the required depth and replaced
with uncontaminated and nutrient balanced soil. This method is practiced in most
parts of India over decades as a remedial measure before sowing crops. Leaching of
F from soil profile is another commonly employed method in India to remove the
contamination below the root zone. This process is essential for poor drainage and
high water table areas that favor the F contamination in the soil. The leaching F
contamination is carried out for the shallow soils using distilled water [143].
Natural remedial measures like crop selection and green manure techniques could
control the soil from being contaminated from F and retain its fertile characteristics.
The crops like Gossypium (Cotton) [144] and Saccharum officinarum (Sugarcane)
are more resistant to the F toxicity. These crops are reported to be capable of
extracting a high amount of F (around 40%) [145] from the soil. For the barren
lands of high F contamination, planting P. juliflora and Nerium oleander plants
could be the effective solution to remove the excess F contamination from the soil
rather than leaving them barren. Such plants could withstand even under extreme F
contamination, and it has also got good market value for its products. Land prepa-
ration with the green manure method has a higher potential to enrich the organic
matter content, thereby promoting microbial activity in the soil. The leguminous
family plants like Sesbania bispinosa, Corchorus (Jute crops), Tephrosia purpurea,
Pithecellobium dulce, Sesbania grandiflora (Agathi keerai), and Vigna aconitifolia
are commonly used as green manure in India. However, the leguminous green
manure is more effective than the plants under the non-leguminous family [146],
subjected to decomposition rate.
There are many in-situ bioremediation techniques like increasing fungi, bacteria,
and other soil organisms that are being adopted by various countries to alleviate the
ill effects of microplastics on the agricultural soil [105, 147, 148]. However, in India,
248 S. Periasamy and R. S. Shanmugam

while the Government of India regulations are framed to address the issue of plastics,
the strict measures to alleviate the effects of microplastics in agriculture are inade-
quate due to the lack of research studies in this domain. Since this is an emerging
serious pollutant that would threaten agricultural production over time, the research
on the effect of microplastics in all agricultural elements deserves special attention in
India.

7.6 Slash-and-Burn Agriculture

The Ministry of Agriculture launched a program “National Policy for Management


of Crop Residue (NPMCR)” in 2014 with the major goals like (1) promotion of
technologies for the effective utilization and in-situ management of crop residue to
prevent the soil from nutrients and minerals loss, (2) promotion of the diversified use
of crop residue for various purposes like industrial raw material, power generation,
charcoal gasification, packing material, etc., (3) creating awareness about the ill
effects of crop burning process and the effective management of crop residues
through the capacity building initiatives for stakeholders, and (4) formulation poli-
cies to control the crop residue burning [90]. Further, stringent measures have been
taken in Rajasthan, Uttar Pradesh, Haryana, and Punjab to reduce the crop residue
burning under National Green Tribunal (NGT) Act (2010). In addition to that, the
Government of India has launched several laws to curb crop residue burning, such as
(1) Section 144 of the Civil Procedure Code (CPC) to ban burning of paddy, (2) The
Air Prevention and Control of Pollution Act (1981), (3) The Environment Protection
Act (1986), (4) The National Tribunal Act (1995), and (5) The National Environ-
ment Appellate Authority Act (1997) [91, 149, 150].

8 Concluding Remarks

Nearly 96.4 Mha of the country’s land area requires the utmost attention to imple-
ment appropriate land management policies to mitigate the ill effects of various
prevailing and emerging degradation processes. Because of the unexpected decline
in agricultural GDP in the late 1990s, India has given its complete attention to
reforming the agriculture development policies. With no significant increase in the
land degradation area from 2005 to 2013, India is very close to achieving land
degradation neutrality status by 2030 as committed. However, reclaiming the
degraded land is equally as important as controlling the degradation rate to increase
the food production by 45% in 2050. Though India is a country with an excellent
record of successful implementation of various broad-scale land-use policies to
mitigate the land degradation processes and increase crop productivity, it still
lacks integration of the existing policies with micro-farm management. The com-
bined execution of conservation agriculture and integrated watershed approach
Agricultural Land Degradation in India 249

requires special attention to combating the agricultural land degradation in the fragile
environment. The land-use policies pertaining to the micro-environment are an
urgent requirement to address multidimensional issues like erosion, salinity, acidity,
and waterlogging in southern peninsular India. Creating awareness about the feasible
alternatives of crop residue burning like bio-gas, bio-char, in-situ treatment, and
economic returns through community programs could significantly reduce the crop
residue burning activity in India. A well-integrated land-use policy for rural fuel-
wood and fodder grazing is needed for the local community to have an economic
balance and reduce the overgrazing practice.
Most importantly, with an increasing population rate, industrial growth, and
erratic climatic behavior, land degradation is expected to increase due to the emerg-
ing pollutants like microplastics and soil fluoride in the future, which is being treated
with the least attention in the present time. So, there is an urgent need for the
government to lay down the new agro-environmental policy and regulations to
curb microplastics’ effects on the agricultural soil. The use of biodegradable mulches
is one of the best solutions to ameliorate the problem. It is also advised to create
awareness about its advantage over plastic mulches through the community pro-
gram. The integrated natural soil mitigation policy is necessitated for the fluoride-
affected parts of India. The policy needs to integrate the natural remedial measures
such as the selection of F-resistant crops, soil replacement, green manure, and crop
residue mulching to prevent the soil from being polluted from F’s upward migration
from the groundwater and increase crop production. However, the supply of F to the
soil through anthropogenic channels like industrial and irrigation activity is unavoid-
able in the semi-arid regions of India. The F-contaminated irrigation water signifi-
cantly controls the microbial activity in the soil. The supply of vermicompost to the
soil is recommended to address this issue, along with other reclamation measures.
The vermicompost supplies all essential nutrients for plant growth and increases
microbial activity in the soil. The use of other inorganic fertilizers should be avoided
for the land patches treated with vermicompost. It is also recommended to treat the
contaminated groundwater before irrigation. Besides, the systematic way of crop
rotation practice with F-absorbing crops also prevents the soil from being contam-
inated continuously.

Acknowledgment The authors acknowledge Ms. Kokila Priya Ravi, Research Scholar, SRM
Institute of Science and Technology, for helping in the visualization work for this chapter. We
extend our sincere gratitude to all the authors we referred to in the chapter. We also thank SRM
Institute of Science and Technology and the Institute of Remote Sensing for providing us with
excellent on-line and off-line accessibility of the reading resources.

References

1. Chand R (2008) India’s agricultural challenges and their implications for growth of its
economy. In: Fujita M (ed) Economic integration in Asia and India. IDE-JETRO series.
Palgrave Macmillan, London. https://doi.org/10.1057/9780230591004_6
250 S. Periasamy and R. S. Shanmugam

2. Census of India (2011) Economic activity. Office of the Registrar General & Census Com-
missioner, New Delhi
3. Minhas PS, Obi Reddy GP (2017) Edaphic stresses and agricultural sustainability: an Indian
perspective. Agric Res 6(1):8–21. https://doi.org/10.1007/s40003-016-0236-4
4. Patel SK, Sharma A, Singh GS (2020) Traditional agricultural practices in India: an approach
for environmental sustainability and food security. Energy Ecol Environ 5(4):253–271. https://
doi.org/10.1007/s40974-020-00158-2
5. Bhattacharyya R, Ghosh BN, Mishra PK, Mandal B, Rao CS, Sarkar D, Das K, Anil KS,
Lalitha M, Hati KM, Franzluebbers AJ (2015) Soil degradation in India: challenges and
potential solutions. Sustainability 7(4):3528–3570. https://doi.org/10.3390/su7043528
6. Maji AK, Reddy GPO, Sarkar D (2010) Prasad R, Pathak PS (eds) Degraded and wasteland of
India. Indian Council of Agricultural Research, New Delhi
7. Ahmad T, Haneef R (2019) Government agricultural schemes in India: a review. Indian
Farmers Digest 52(11):20–24
8. NBSS & LUP (2015) Annual report. National Bureau of Soil Survey and Land Use Planning
(NBSS & LUP), Nagpur
9. Mythili G, Goedecke J (2016) Economics of land degradation in India. In: Nkonya E,
Mirzabaev A, von Braun J (eds) Economics of land degradation and improvement – a global
assessment for sustainable development. Springer, Cham. https://doi.org/10.1007/978-3-319-
19168-3_15
10. Gore SR, Bhagwat KA (1991) Saline degradation of Indian agricultural lands: a case study in
Khambhat Taluka, Gujarat state (India), using satellite remote sensing. Geocarto Int 6(3):5–13.
https://doi.org/10.1080/10106049109354315
11. Qadir M, Quillérou E, Nangia V, Murtaza G, Singh M, Thomas RJ, Drechsel P, Noble AD
(2014) Economics of salt-induced land degradation and restoration. Nat Res Forum 38(4):
282–295. https://doi.org/10.1111/1477-8947.12054
12. Mishra B, Ramakrishnan P (1983) Slash and burn agriculture at higher elevations in north-
eastern India. II. Soil fertility changes. Agric Ecosyst Environ 9(1):83–96. https://doi.org/10.
1016/0167-8809(83)90008-7
13. Jalees K, Vemuri R (1980) Pesticide pollution in India. Int J Environ Stud 15(1):49–53. https://
doi.org/10.1080/00207238008737423
14. Veerasingam S, Ranjani M, Venkatachalapathy R, Bagaev A, Mukhanov V, Litvinyuk D,
Verzhevskaia L, Guganathan L, Vethamony P (2020) Microplastics in different environmental
compartments in India: analytical methods, distribution, associated contaminants and research
needs. TrAC – Trends Analyt Chem 133:1–13. https://doi.org/10.1016/j.trac.2020.116071
15. Cronin SJ, Manoharan V, Hedley MJ, Loganathan P (2000) Fluoride: a review of its fate,
bioavailability, and risks of fluorosis in grazed-pasture systems in New Zealand. N Z J Agric
Res 43(3):295–321. https://doi.org/10.1080/00288233.2000.9513430
16. Bhattacharya P, Samal AC (2018) Fluoride contamination in groundwater, soil and cultivated
foodstuffs of India and its associated health risks: a review. Res J Recent Sci 7(4):36–47
17. Shrivastava A, Ghosh D, Dash A, Bose S (2015) Arsenic contamination in soil and sediment in
India: sources, effects, and remediation. Curr Pollut Rep 1:35–46. https://doi.org/10.1007/
s40726-015-0004-2
18. Farooq SH, Chandrasekharam D, Dhanachandra W, Ram K (2019) Relationship of arsenic
accumulation with irrigation practices and crop type in agriculture soils of Bengal Delta, India.
Appl Water Sci 9(119):1–11. https://doi.org/10.1007/s13201-019-0904-1
19. IPCC (2007) Fourth assessment report: climate change – a synthesis report. An assessment of
the Intergovernmental Panel on Climate Change (IPCC) at Valencia, Spain
20. NBSS & LUP (2012) Annual report. National Bureau of Soil Survey and Land Use Planning
(NBSS & LUP), Nagpur
21. Singh SK, Batta RK, Chatterji S (2016) Land use planning for arresting land degradation,
combating climate change and ensuring food security – a training manual. ICAR-National
Bureau of Soil Survey and Land Use Planning Publication 171, Nagpur, p 135
Agricultural Land Degradation in India 251

22. United Nations Sustainable Development (2012) The future we want. Resolution adopted by
the United Nations General Assembly, sixty-sixth session, Agenda item 19
23. Population Division (2015) United Nations Department of economic and social affairs. http://
www.un.org/en/development/desa/news/population/2015-report.html
24. Hatti N (1977) Impact of assistance under P.L. 480 on Indian economy 1956–1970. Econ Hist
20(1):23–40. https://doi.org/10.1080/00708852.1977.11877475
25. Sekhar CSC (2014) Indian agriculture – a review of policy and performance. Yojana 6:32–36
26. Srivastava P, Balhara M, Giri B (2020) Soil health in India: past history and future
perspective. In: Giri B, Varma A (eds) Soil health. Soil biology, vol 59. Springer, Cham.
https://doi.org/10.1007/978-3-030-44364-1_1
27. Nethrayini KR, Mundinamani SM (2013) Impact of technology mission on oilseeds and pulses
on pulse production in Karnataka. Int Res J Agric Econ Statist 4(2):148–153
28. Arora VPS (2013) Agricultural policies in India: retrospect and prospect. Agric Econ Res Rev
26(2):135–157
29. Saini S, Gulati A (2016) India’s food security policies in the wake of global food price
volatility. In: Kalkuhl M, von Braun J, Torero M (eds) Food price volatility and its implica-
tions for food security and policy. Springer, Cham. https://doi.org/10.1007/978-3-319-28201-
5_14
30. Behera B, Mishra P (2007) Acceleration of agricultural growth in India: suggestive policy
framework. Econ Polit Wkly 42(42):4268–4271
31. Sharda VN, Dogra P, Prakash C (2010) Assessment of production losses due to water erosion
in rainfed areas of India. J Soil Water Conserv 65(2):79–91. https://doi.org/10.2489/jswc.65.
2.79
32. Reddy VR (2003) Land degradation in India: extent, costs and determinants. Econ Polit Wkly
38(44):4700–4713. https://doi.org/10.2307/4414225
33. The Energy and Researches Institute (2018) Economics of desertification, land degradation
and drought in india, vol I: macroeconomic assessment of the costs of land degradation in
India. Ministry of Environment, Forest and Climate Change, Government of India
34. NCA (1976) Report of the National Commission on agriculture. National Commission of
Agriculture (NCA), Government of India, New Delhi, pp 427–472
35. MoA (1978) Indian agriculture in brief.17th edn. Directorate of Economics and Statistics,
Ministry of Agriculture (MoA), Department of Agriculture and Cooperation, Ministry of
Agriculture and irrigation, Government of India, New Delhi
36. Vohra BB (1980) A policy for land and water, vol 18. Department of Environment, Govern-
ment of India, New Delhi, pp 64–70
37. Bhumbla DR, Khare A (1984) Estimate of wastelands in India. Society for Promotion of
Wastelands Development, Allied, New Delhi, p 18
38. MoA (1985) Indian agriculture in brief.20th edn. Directorate of Economics and Statistics,
Ministry of Agriculture (MoA), Department of Agriculture and Cooperation, Ministry of
Agriculture, Government of India, New Delhi
39. NBSS & LUP (1994) Global assessment of soil degradation guidelines. National Bureau of
Soil Survey and Land Use Planning (NBSS & LUP), Nagpur
40. NBSS & LUP (2004) Annual report. National Bureau of Soil Survey and Land Use Planning
(NBSS & LUP), Nagpur
41. Gautam NC, Narayan LRA (1988) Wastelands in India. Pink Publishing House, Mathura, p 96
42. Nawar S, Buddenbaum H, Hill J (2015) Digital mapping of soil properties using multivariate
statistical analysis and ASTER data in an Arid region. Remote Sens 7(2):1181–1205. https://
doi.org/10.3390/rs70201181
43. Rao DP (1999) Remote sensing application for land use and urban planning: retrospective and
perspective. In: Proceedings of the ISRS national symposium on remote sensing application
for natural resources retrospective and perspective, Bangalore, pp 287–297
44. NRSA (2000) Waste land atlas of India. Government of India, Hyderabad
252 S. Periasamy and R. S. Shanmugam

45. Breunig FM, Galvão LS, Formaggio AR (2008) Detection of sandy soil surfaces using
ASTER-derived reflectance, emissivity and elevation data: potential for the identification of
land degradation. Int J Remote Sens 29(6):1833–1840. https://doi.org/10.1080/
01431160701851791
46. SAC (2016) Desertification and land degradation Atlas of India (Based on IRS AWiFS data of
2011–13 and 2003–05). Space Applications Centre (SAC), Indian Space Research Organisa-
tion (ISRO), Government of India, Ahmedabad, pp 1–219
47. NBSS&LUP (2015) Annual report 2014–15. National Bureau of Soil Survey and Land Use
Planning, Nagpur
48. Borah B, Bhattacharjee A, Ishwar NM (2018) Bonn challenge and India: progress on resto-
ration efforts across states and landscapes, vol viii. IUCN and MoEFCC, Government of India,
New Delhi, p 32. https://doi.org/10.2305/IUCN.CH.2018.12.en
49. Behera UK, France J (2016) Integrated farming systems and the livelihood security of small
and marginal farmers in India and other developing countries. Adv Agron 138:235–282.
https://doi.org/10.1016/bs.agron.2016.04.001
50. Dolma K, Rishi MS, Lata R (2020) State of groundwater resource: relationship between its
depth and sewage contamination in Leh town of Union Territory of Ladakh. Appl Water Sci
10(3):1–18. https://doi.org/10.1007/s13201-020-1157-8
51. Santra P, Moharana PC, Kumar M, Soni ML, Pandey CB, Chaudhari SK, Sikka AK (2017)
Crop production and economic loss due to wind erosion in hot arid ecosystem of India.
Aeolian Res 28:71–82. https://doi.org/10.1016/j.aeolia.2017.07.009
52. Kar A, Moharana PC, Raina P, Kumar M, Soni ML, Santra P, Ajai AAS, Dhinwa PS (2009)
Desertification and its control measures. In: Kar A, Garg BK, Singh MP, Kathju S (eds) Trends
in arid zone research in India. Central Arid Zone Research Institute, Jodhpur, pp 1–47
53. Kalsi R (2007) Status, distribution and management of Galiformes in arid and semi-arid zones
of India. In: Sathyakumar S, Sivakumar K (eds) Galliformes of India. ENVIS bulletin: wildlife
and protected areas, vol 10. Wildlife Institute of India, Dehradun, p 252
54. Zorn M, Komac B (2013) Erosion. In: Bobrowsky PT (ed) Encyclopedia of natural hazards.
Encyclopedia of earth sciences series. Springer, Dordrecht. https://doi.org/10.1007/978-1-
4020-4399-4_120
55. Sharma KD (1997) Assessing the impact of overgrazing on soil erosion in arid regions at a
range of spatial scales. Human impact on erosion and sedimentation. In: Proceedings of the
international symposium of the fifth scientific assembly of the International Association of
Hydrological Sciences (IAHS). Rabat, pp 119–123
56. Zachar D (1982) Soil erosion. Developments in soil science, vol 10. Elsevier, Amsterdam,
p 547
57. Seitz S, Goebes P, Puerta VL, Pereira EIP, Wittwer R, Six J, van der Heijden MGA, Scholten
T (2019) Conservation tillage and organic farming reduce soil erosion. Agron Sustain Dev
39(4):1–10. https://doi.org/10.1007/s13593-018-0545-z
58. Dutta D, Das S, Kundu A, Taj A (2015) Soil erosion risk assessment in Sanjal watershed,
Jharkhand (India) using geo-informatics, RUSLE model and TRMM data. Model Earth Syst
Environ 1(37):1–10. https://doi.org/10.1007/s40808-015-0034-1
59. Mandal D, Sharda VN (2011) Assessment of permissible soil loss in India employing a
quantitative bio-physical model. Curr Sci 100(3):383–390
60. Pimentel D (2006) Soil erosion: a food and environmental threat. Environ Dev Sustain 8(1):
119–137. https://doi.org/10.1007/s10668-005-1262-8
61. Mertia RS, Santra P, Kandpal BK, Prasad R (2010) Mass-height profile and total mass
transport of wind eroded aeolian sediments from rangelands of Indian Thar desert. Aeolian
Res 2(2–3):135–142. https://doi.org/10.1016/j.aeolia.2010.04.002
62. Nachshon U (2018) Cropland soil salinization and associated hydrology: trends, processes and
examples. Water 10(8):1–20. https://doi.org/10.3390/w10081030
63. Kumar P, Sharma PK (2020) Soil salinity and food security in India. Front Sustain Food Syst
4:1–15. https://doi.org/10.3389/fsufs.2020.533781
Agricultural Land Degradation in India 253

64. CSSRI (2015) Vision 2050. ICAR-Central Soil Salinity Research Institute (CSSRI), Haryana
65. Dagar JC (2005) Salinity research in India: an overview. Bull Natl Inst Ecol 15:69–80
66. Shahid SA, Zaman M, Heng L (2018) Soil salinity: historical perspectives and a world
overview of the problem. In: Guideline for salinity assessment, mitigation and adaptation
using nuclear and related techniques. Springer, Cham, pp 43–53. https://doi.org/10.1007/978-
3-319-96190-3_2
67. Mandal S, Raju R, Kumar A, Kumar P, Sharma PC (2018) Current status of research,
technology response and policy needs of salt-affected soils in India – a review. J Ind Soc
Coast Agric Res 36(2):40–53
68. Mitch JW, Hernandez ME (2013) Landscape and climate change threats to wetlands of North
and Central America. Aquat Sci 75:133–149. https://doi.org/10.1007/s00027-012-0262-7
69. Nachshon U, Ireson A, van der Kamp G, Davies SR, Wheater HS (2014) Impacts of climate
variability on wetland salinization in the North American prairies. Hydrol Earth Syst Sci 18(4):
1251–1263. https://doi.org/10.5194/hess-18-1251-2014
70. Rengasamy P (2006) World salinization with emphasis on Australia. J Exp Bot 57(5):
1017–1023. https://doi.org/10.1093/jxb/erj108
71. CWC (2017) Problems of salination of land in coastal areas of India and suitable protection
measures. Hydrological Studies Organization Central Water Commission, Government of
India, pp 1–345
72. Sharma SS, Sharma RP, Singh RS, Singh SK (2016) Characterization and formation of salt
affected soils in Bhilwara district, Rajasthan. Agropedology 26(02):230–236
73. Kolarkar AS, Dhir RP, Singh N (1980) Characteristics and morphogenesis of salt-affected
soils in southeastern arid Rajasthan. J Ind Soc Photo-interpret Remote Sens 8:31–41. https://
doi.org/10.1007/BF02990666
74. GWD (2017) Ground water level scenario in Rajasthan, pre and post monsoon survey.
Government of Rajasthan, Ground Water Department (GWD), pp 1–206
75. IDNP (2002) Recommendations on waterlogging and salinity control based on pilot area
drainage research. Indo-Dutch Network Project (IDNP), Central Soil Salinity Research Insti-
tute, Karnal, p 100
76. ICID (2003) Important data of ICID member countries. International Commission on Irriga-
tion and Drainage (ICID). http://www.icid.org
77. Ritzema HP, Satyanarayana TV, Raman S, Boonstra J (2008) Subsurface drainage to combat
waterlogging and salinity in irrigated lands in India: lessons learned in farmers’ fields. Agric
Water Manag 95(3):179–189. https://doi.org/10.1016/j.agwat.2007.09.012
78. Dattaa KK, de Jong C, Singha OP (2000) Reclaiming salt-affected land through drainage in
Haryana, India: a financial analysis. Agric Water Manag 46(1):55–71. https://doi.org/10.1016/
S0378-3774(00)00077-9
79. Masilamani P, Arulmozhiselvan K, Alagesan A (2020) Prospects of biodrainage to mitigate
problems of waterlogging and soil salinity in context of India – a review. J Appl Nat Sci 12(2):
229–243. https://doi.org/10.31018/jans.vi.2285
80. Joshi PK, Jha D (1992) An economic inquiry into the impact of soil alkalinity and
waterlogging. Ind J Agric Econ 47(2):196–204
81. Pandey AC, Singh SK, Nathawat MS (2010) Waterlogging and flood hazards vulnerability and
risk assessment in Indo Gangetic plain. Nat Hazards 55(2):273–289. https://doi.org/10.1007/
s11069-010-9525-6
82. Singh SK, Kanga S (2017) Mapping of salt affected and waterlogged areas using geospatial
technique. Int J Recent Innov Trends Comput Commun 5(5):1298–1305
83. Gain AK, Benson D, Rahman R, Datta DK, Rouillard JJ (2017) Tidal river management in the
south west Ganges-Brahmaputra delta in Bangladesh: moving towards a transdisciplinary
approach? Environ Sci Pol 75:111–120
84. Goyal VC, Jain SK, Pareek N (2005) Water logging and drainage assessment in Ravi-Tawi
irrigation command (J&K) using remote sensing approach. J Ind Soc Remote Sens 33(1):7–15.
https://doi.org/10.1007/BF02989986
254 S. Periasamy and R. S. Shanmugam

85. (1976) Report of the national commission on agriculture, part V, IX and abridged report.
Ministry of Agriculture and Irrigation, Government of India, New Delhi
86. Sharma VP, Thaker H (2011) Demand for fertiliser in India: determinants and outlook for
2020. IIMA working papers WP2011-04-01. Indian Institute of Management Ahmedabad,
Research and Publication Department
87. Ministry of Agriculture and Farmers Welfare (2015–2016) Impact of chemical fertilizers and
pesticides on agriculture and allied sectors in the country, twenty-ninth report. Standing
Committee on Agriculture, Ministry of Agriculture and Farmers Welfare, Government of
India, pp 1–157
88. Kumar S, Singh P (2010) Inclusive agricultural growth: district-wise agricultural productivity
analysis in Punjab. In: Kannan E, Lokesh GB (eds) Proceedings of national workshop on
inclusive agricultural growth: regional perspective, pp 21–51
89. Patra S, Mishra P, Mahapatra SC, Mithun SK (2016) Modelling impacts of chemical fertilizer
on agricultural production: a case study on Hooghly district, West Bengal, India. Model Earth
Syst Environ 2(4):1–11. https://doi.org/10.1007/s40808-016-0223-6
90. NPMCR (2014) Incorporation in soil and mulching baling/binder for domestic/industrial as
fuel. National Policy for Management of Crop Residues (NPMCR), Government of India,
Ministry of Agriculture Department of Agriculture & Cooperation. http://agricoop.nic.in/sites/
default/files/NPMCR_1.pdf. Accessed 10 Mar 2021
91. Bhuvaneshwari S, Hettiarachchi H, Meegoda JN (2019) Crop residue burning in India: policy
challenges and potential solutions. Int J Environ Res Public Health 16(5):1–19. https://doi.org/
10.3390/ijerph16050832
92. Devi S, Gupta C, Jat SL, Parmar MS (2017) Crop residue recycling for economic and
environmental sustainability: the case of India. Open Agric 2:486–494. https://doi.org/10.
1515/opag-2017-0053
93. Jain N, Bhatia A, Pathak H (2014) Emission of air pollutants from crop residue burning in
India. Aerosol Air Qual Res 14(1):422–430. https://doi.org/10.4209/aaqr.2013.01.0031
94. Meshram JR (2002) Biomass resources assessment programme and prospects of biomass as an
energy resource in India. IREDA News 13(4):21–29
95. MNRE (2009) Annual report of the Ministry of New and Renewable Energy (MNRE).
Government of India, New Delhi
96. Pathak H, Bhatia A, Jain N, Aggarwal PK (2010) Greenhouse gas emission and mitigation in
Indian agriculture – a review. In: Singh B (ed) ING bulletins on regional assessment of reactive
nitrogen, bulletin no. 19. SCON-ING, New Delhi, p 34
97. Sarkar S, Skalicky M, Hossain A, Brestic M, Saha S, Garai S, Ray K, Brahmachari K (2020)
Management of crop residues for improving input use efficiency and agricultural sustainabil-
ity. Sustainability 12(23):1–24. https://doi.org/10.3390/su12239808
98. Abdurrahman MI, Chaki S, Saini G (2020) Stubble burning: effects on health & environment,
regulations and management practices. Environ Adv 2:1–12. https://doi.org/10.1016/j.envadv.
2020.100011
99. Anbumani S, Kakkar P (2018) Ecotoxicological effects of microplastics on biota: a review.
Environ Sci Pollut Res 25(15):14373–14396. https://doi.org/10.1007/s11356-018-1999-x
100. Microplastics in Agriculture – ‘Not’ a Micro Issue (2020) Agripedia. https://krishijagran.com/
agripedia/microplastics-in-agriculture-not-a-micro-issue/
101. Karthik R, Robin RS, Purvaja R, Ganguly D, Anandavelu I, Raghuraman R, Hariharan G,
Ramakrishna A, Ramesh R (2018) Microplastics along the beaches of southeast coast of India.
Sci Total Environ 645:1388–1399. https://doi.org/10.1016/j.scitotenv.2018.07.242
102. Madhav NV, Gopinath KP, Krishnan A, Rajendran N, Krishnan A (2020) A critical review on
various trophic transfer routes of microplastics in the context of the Indian coastal ecosystem.
Watershed Ecol Environ 2:25–41. https://doi.org/10.1016/j.wsee.2020.08.001
103. Sarker A, Deepo DM, Nandi R, Rana J, Islam S, Rahman S, Hossain MN, Islam MS, Baroi A,
Kim J (2020) A review of microplastics pollution in the soil and terrestrial ecosystems: a
Agricultural Land Degradation in India 255

global and Bangladesh perspective. Sci Total Environ 733:1–14. https://doi.org/10.1016/j.


scitotenv.2020.139296
104. de Souza Machado AA, Lau CW, Kloas W, Bergmann J, Bachelier JB, Faltin E, Becker R,
Görlich AS, Rillig MC (2019) Microplastics can change soil properties and affect plant
performance. Environ Sci Technol 53(10):6044–6052. https://doi.org/10.1021/acs.est.
9b01339
105. Pathan SI, Arfaioli P, Bardelli T, Ceccherini MT, Nannipieri P, Pietramellara G (2020) Soil
pollution from micro- and nanoplastic debris: a hidden and unknown biohazard. Sustainability
12(18):1–31. https://doi.org/10.3390/su12187255
106. Singh G, Sinam G, Kriti PM, Kumari B, Kulsoom M (2020) Soil pollution by fluoride in India:
distribution, chemistry and analytical methods. In: Shukla V, Kumar N (eds) Environmental
concerns and sustainable development. Springer, Singapore. https://doi.org/10.1007/978-981-
13-6358-0_12
107. Chatterjee A, Sarah S, Sreedevi PD, Selles A, Ahmed S (2017) Demarcation of fluoride
vulnerability zones in granitic aquifer, semi-arid region, Telengana, India. Arab J Geosci
10(558):1–18. https://doi.org/10.1007/s12517-017-3334-0
108. Reddy SKK, Sahadevan DK, Gupta H, Reddy DV (2019) GIS-based prediction of ground-
water fluoride contamination zones in Telangana, India. J Earth Syst Sci 128(132):1–18.
https://doi.org/10.1007/s12040-019-1151-4
109. Teotia SP, Teotia M (1984) Endemic fluorosis in India: a challenging national health problem.
J Assoc Physicians India 32(4):347–352
110. Singh G, Kumari B, Sinam G, Kriti KN, Mallick S (2018) Fluoride distribution and contam-
ination in the water, soil and plants continuum and its remedial technologies, an Indian
perspective – a review. Environ Pollut 239:95–108. https://doi.org/10.1016/j.envpol.2018.
04.002
111. Rai K, Agarwal M, Dass S, Shrivastav R (2000) Fluoride: diffusive mobility in soil and some
remedial measures to control its plant uptake. Curr Sci 79(9):1370–1373. http://www.jstor.org/
stable/24105289
112. Chatterjee N, Sahu G, Bag AG, Pal B, Hazra GC (2020) Role of fluoride on soil, plant and
human health: a review on its sources, toxicity and mitigation strategies. Int J Environ Climate
Change 10(8):77–90. https://doi.org/10.9734/IJECC/2020/v10i830220
113. Mishra PC, Meher K, Bhosagar D, Pradhan K (2009) Fluoride distribution in different
environmental segments at Hirakud Orissa (India). Afr J Environ Sci Technol 3(9):260–264.
https://doi.org/10.5897/AJEST09.077
114. Yadav N, Rani K, Yadav SS, Yadav DK, Yadav VK, Yadav N (2018) Soil and water pollution
with fluoride, geochemistry, food safety issues and reclamation – a review. Int J Curr
Microbiol App Sci 7(5):1147–1162. https://doi.org/10.20546/ijcmas.2018.705.140
115. Dehbandi R, Moore F, Keshavarzi B (2018) Geochemical sources, hydrogeochemical behav-
ior, and health risk assessment of fluoride in an endemic fluorosis area, central Iran.
Chemosphere 193:763–776. https://doi.org/10.1016/j.chemosphere.2017.11.021
116. Sankhla MS, Kumar R (2018) Fluoride contamination of water in India and its impact on
public health. ARC J Forens Sci 3(2):10–15. https://doi.org/10.20431/2456-0049.0302002
117. Chaudhary V, Satheeshkumar S (2018) Assessment of groundwater quality for drinking and
irrigation purposes in arid areas of Rajasthan, India. Appl Water Sci 8(218):1–17. https://doi.
org/10.1007/s13201-018-0865-9
118. Haidouti C, Chronopoulou A, Chronopoulos J (1993) Effects of fluoride emissions from
industry on the fluoride concentration of soils and vegetation. Biochem Syst Ecol 21(2):
195–208. https://doi.org/10.1016/0305-1978(93)90037-R
119. Gago C, Romar A, Fernandez-Marcos ML, Alvarez E (2012) Fluorine sorption by soils
developed from various parent materials in Galicia (NW Spain). J Colloid Interface Sci
374(1):232–236. https://doi.org/10.1016/j.jcis.2012.01.047
256 S. Periasamy and R. S. Shanmugam

120. Kumar K, Giri A, Vivek P, Kalaiyarasan T, Kumar B (2017) Effects of fluoride on respiration
and photosynthesis in plants: an overview. Peertechz J Environ Sci Toxicol 2(1):43–47.
https://doi.org/10.17352/aest.000011
121. Mandal D, Sharda VN (2011) Appraisal of soil erosion risk in the eastern Himalayan region of
India for soil conservation planning. Land Degrad Dev 24(5):430–437. https://doi.org/10.
1002/ldr.1139
122. Pathak P, Mishra PK, Rao KV, Wani SP, Sudi R (2009) Best options on soil and water
conservation. In: Best bet options for integrated watershed management, proceedings of the
comprehensive assessment of watershed programs in India, Andhra Pradesh
123. Saha R, Tomar JMS, Ghosh PK (2007) Evaluation and selection of multi-purpose tree for
improving soil hydro-physical behaviour under hilly eco-system of north east India. Agrofor
Syst 69:239–247. https://doi.org/10.1007/s10457-007-9044-y
124. Bhattacharyya R, Tuti MD, Kundu S, Bisht JK, Bhatt JC (2012) Conservation tillage impacts
on soil aggregation and carbon pools in a sandy clay loam soil of the Indian Himalayas. Soil
Sci Soc Am J 76(2):617–627. https://doi.org/10.2136/sssaj2011.0320
125. Sharma PD, Sarkar AK (2005) Managing acid soils for enchancing productivity. Technical
Bulletin, NRM Division, KAB-II, New Delhi, p 23
126. Bhat JA, Mandal B, Hazra GC (2007) Basic slag as a liming material to ameliorate soil acidity
in Alfisols of sub-tropical India. Am-Euras J Agric Environ Sci 2(4):321–327
127. Sekar I, Pal S (2012) Rice and wheat crop productivity in the indo-gangetic plains of India:
changing pattern of growth and future strategies. Ind J Agric Econ 67(2):238–252. https://doi.
org/10.22004/ag.econ.204809
128. Bhattacharyya R, Ghosh BN, Dogra P, Mishra PK, Santra P, Kumar S, Fullen MA, Mandal
UK, Anil KS, Lalitha M, Sarkar D, Mukhopadhyay D, Das K, Pal M, Yadav R, Chaudhary
VP, Parmar B (2016) Soil conservation issues in India. Sustainability 8(6):1–37. https://doi.
org/10.3390/su8060565
129. Mandal B, Majumder B, Bandopadhyay PK, Hazra GC, Gangopadhyay A, Samantaroy RN,
Misra AK, Chowdhuri J, Saha MN, Kundu S (2007) The potential of cropping systems and
soil amendments for carbon sequestration in soils under long-term experiments in subtropical
India. Glob Chang Biol 13(2):357–369. https://doi.org/10.1111/j.1365-2486.2006.01309.x
130. Ghosh BN, Dogra P, Bhattacharyya R, Sharma NK, Dadhwal KS (2012) Effects of grass
vegetative strips on soil conservation and crop yield under rainfed conditions in the Indian
sub-Himalayas. Soil Use Manag 28(4):635–646. https://doi.org/10.1111/j.1475-2743.2012.
00454.x
131. Bhimaya CP, Kaul RN, Ganguli BN (1960) Sand dune rehabilitation in Western Rajasthan. In:
Proceedings of the fifth world forestry congress, Seattle, pp 358–363
132. Bhimaya CP, Choudhary MD (1961) Plantation of windbreaks in the central mechanized farm,
Suratgarh. Indian Forester 87(6):354–367
133. Gupta JP, Rao GGSN, Gupta GN, Ramana Rao BV (1983) Soil drying and wind erosion as
affected by different types of shelterbelts planted in the desert region of Western Rajasthan,
India. J Arid Environ 6(1):53–58. https://doi.org/10.1016/S0140-1963(18)31432-0
134. Mann HS (1985) Wind erosion and its control. FAO conservation guide no. 10. FAO, Rome
135. Fryrear DW, Bilbro JD (1994) Wind erosion control with residues and related practices. In:
Unger PW (ed) Managing agricultural residues. Lewis, Boca Raton, pp 7–17
136. Yadav JSP (1989) Irrigation induced soil salinity and sodicity. In: Proceedings of the world
food day symposium on environmental problems affecting agriculture in the Asia and Pacific
region, Bangkok, pp 47–62
137. Rao KVGK, Biswas CR, Bandyopadhyay AK (1981) Developing suitable methods to store
rain water and its use for irrigation in winter. Annual report. Central Soil Salinity Research
Institute, Karnal, pp 190–191
138. Burman D, Bandyopadhyay BK, Mandal S, Mandal UK, Mahanta KK, Sarangi SK, Maji B,
Rout S, Bal AR, Gupta SK, Sharma DK (2013) Land shaping – a unique technology for
Agricultural Land Degradation in India 257

improving productivity of coastal land. Bulletin no. CSSRI/Canning Town/Bulletin/2013/02.


Central Soil Salinity Research Institute, Regional Research Station, Canning Town, p 38
139. CSSRI, NAIP (2014) Final report of NAIP sub-project on: strategies for sustainable manage-
ment of degraded coastal land and water for enhancing livelihood security of farming
communities (component 3, GEF funded). In: Burnan D, Mandal S, Mahanta KK (eds).
Central Soil Salinity Research Institute (CSSRI), Regional Research Station (RRS), Canning
Town, p 104
140. Singh P, Aggarwal PK, Bhatia VS, Murty MVR, Pala M, Owens T, Benli B, Rao KPC, Wani
SP (2009) Yield gap analysis: modeling of achievable yields at farm level. In: Wani SP, Wani
SP, Rockström J, Oweis T (eds) Rainfed agriculture: unlocking the potential. CAB Int.,
Wallingford, pp 81–123. https://doi.org/10.1079/9781845933890.0081
141. Joseph S, Manoj M (1989) Conservation measures for agricultural lands. Ind J Soil Conserv
17:1–5
142. Natarajan A, Janakiraman M, Manoharan S, AKS K, Vadivelu S, Sarkar D (2010) Assessment
of land degradation and its impact on land resources of Sivagangai block, Tamil Nadu,
India. In: Zdruli P, Pagliai M, Kapur S, Faz Cano A (eds) Land degradation, and desertifica-
tion: assessment, mitigation and remediation. Springer, Berlin, pp 235–252
143. Gao Z, Shi M, Zhang H, Feng J, Fang S, Cui Y (2020) Formation and in situ treatment of high
fluoride concentrations in shallow groundwater of a semi-arid region: Jiaolai Basin, China. Int
J Environ Res Public Health 17(21):1–24. https://doi.org/10.3390/ijerph17218075
144. Kumar S, Singh M (2015) Effect of fluoride contaminated irrigation water on eco-physiology,
biomass and yield in Gossypium hirsutum L. Trop Plant Res 2(2):134–142
145. Santos-Díaz MS, Zamora-Pedrazaa C (2010) Fluoride removal from water by plant species
that are tolerant and highly tolerant to hydrogen fluoride. Fluoride 43(2):150–156
146. Bhattarai N, Vaidya GS, Baral B (2012) Effect of mycorrhizal soil and green manures on
growth of Ipil Ipil (Leucaenadiversifolia L.). Scientific World 10(10):66–69. https://doi.org/
10.3126/sw.v10i10.6864
147. Muenmee S, Chiemchaisri W, Chiemchaisri C (2015) Microbial consortium involving bio-
logical methane oxidation in relation to the biodegradation of waste plastics in a solid waste
disposal open dump site. Int Biodeterior Biodegradation 102:172–181. https://doi.org/10.
1016/j.ibiod.2015.03.015
148. Sekhar VC, Nampoothiri KM, Mohan AJ, Nair NR, Bhaskar T, Pandey A (2016) Microbial
degradation of high impact polystyrene (HIPS), an e-plastic with decabromodiphenyl oxide
and antimony trioxide. J Hazard Mater 318:347–354
149. Kumar P, Kumar S, Joshi L (2015) The extend and management of crop residue stubbles. In:
Kumar P, Kumar S, Joshi L (eds) Socioeconomic and environmental implications of agricul-
tural residue burning: a case study of Punjab, India. Springer, Berlin, pp 13–64. https://doi.org/
10.1007/978-81-322-2014-5_2
150. Lohan SK, Jat HS, Yadav AK, Sidhu HS, Jat ML, Choudhary M, Peter JK, Sharma PC (2018)
Burning issues of paddy residue management in north-west states of India. Renew Sust Energ
Rev 81(1):693–706. https://doi.org/10.1016/j.rser.2017.08.057
Degradation of Agricultural Lands in Israel

Gil Eshel, Elazar Volk, Alon Maor, Eli Argaman, and Guy J. Levy

Contents
1 Israeli Agriculture and Soil Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
2 Soil Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
2.1 Soil Erosion by Water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
2.2 Soil Erosion by Wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
3 Loss of Organic Carbon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
4 Agronomic Approaches for Soil Degradation Mitigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
5 Salinization and Sodification by Irrigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
6 Agriculture Land Loss to Construction and Infrastructure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
7 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270

Abstract Agriculture in Israel is practiced under unique and extreme conditions,


including long hot summers, many cycles of drought years, and shortage of good
quality water. Consequently, extensive use of marginal water, comprising mainly
municipal treated wastewater (TWW) on >60% of the irrigated land. Israel’s soil
degradation arises primarily from extensive agricultural activities such as frequent
tilling, loss of organic carbon, irrigation with variable water quality (i.e., salination
and sodification), the impact of climate change on rain patterns, and soil loss due to
urbanization and infrastructure. According to estimates, ~70% of the agricultural
land in Israel is in danger of soil erosion, of which nearly half is at severe risk, and
about half is at moderate risk. In the regions susceptible to soil erosion, soil loss,
mainly by water, is estimated at a rate of about 2–4 mm per year. Awareness in Israel

G. Eshel (✉), E. Volk, and E. Argaman


Soil Erosion Research Station, Ministry Of Agriculture and Rural Development, Rishon
LeZion, Israel
A. Maor and G. J. Levy
Soil Conservation and Drainage Department, Ministry Of Agriculture and Rural Development,
Rishon LeZion, Israel
Agricultural Research Organization (ARO), Volcani Institute, Rishon LeZion, Israel

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 259
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 259–272, DOI 10.1007/698_2022_931,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 14 December 2022
260 G. Eshel et al.

by both farmers and policymakers of the importance of employing soil conservation


practices to prevent further degradation of the country’s soil resource increased
significantly. Leading conservation practices include no or reduced tillage coupled
with temporal manure application and use of crop rotation and service crops. Yet,
current measures need to be further encouraged and strengthened to ensure sustain-
able use of Israel’s soils so that the needs of future generations can be met.

Keywords Conservation agriculture, Soil organic carbon, Soil salinization and


sodification, Soils, Water erosion, Wind erosion

1 Israeli Agriculture and Soil Degradation

Israel is situated in the eastern part of the Mediterranean basin with wet and cold
winters and hot and dry summers. It stretches along a climatic gradient, ranging from
the sub-humid Mediterranean in the North of the country to hyper-arid in the South
[1]. The soil parent materials are mostly carbonate rocks and eolian dust. The soil
scrapes are associated with highly diverse relief and vegetation [2]. Alfisols,
Mollisols, and Vertisols (Grumosols) are common in Israel’s northern and central
parts, and Aridisols are common in the South and Jordan Valley. Entisols and
Inceptisols, influenced by intense geomorphological processes and anthropogenic
activity, occur all over the country [3].
Israeli agricultural lands occupy an area of about 400,000 ha, which is about 20%
of the state’s land. About 80% of the farming activities occur in Israel’s southern and
northern districts, characterized by arid and semi-arid to semi-wet moist moisture
regimes, respectively. The majority of the agricultural activity is considered very
intensive, with about 45% of the cropped land being under irrigation. Concerning
irrigation water quality, Israeli agriculture currently faces a unique situation with the
recent introduction of large volumes of desalinized water into the country’s water
system. Hence, water allocated to agriculture includes (1) freshwater originating
from aquifers and the Lake of Galilee, desalinized water, and a mixture of the two,
and (2) marginal water comprising mainly municipal treated wastewater (TWW)
whose water of origin vary between freshwater, desalinized water, and a mixture
thereof. Marginal water is used for more than 60% of the irrigated land with an
annual volume of 384 × 106 m3 [4] Agriculture contributes a relatively small portion
to the Israeli economy, consisting of roughly 1% of the total GDP and about 1.8% of
exported goods [5]. Yet, Israeli agriculture is of great importance regarding fresh
produce supply and for fresh and nutritious food security. Israel is known for
growing a wide range of crop types. These include crops ranging from tropical
(Banana, Pineapple), and sub-tropical (Avocado and Mango) up to classical dessert
plantations (palms and olives). Field crops occupy 63% of the agricultural area,
followed by orchards 30%, with vegetables in greenhouses and flowers occupying
3% and 1%, respectively (Fig. 1).
Degradation of Agricultural Lands in Israel 261

Fig. 1 Main categories of Non Cultivated; 10


agricultural land use in Vegtables in green Flowers; 6
Israel; numbers in the pie houses; 11
chart represent the thousand
ha area (data source: Israeli
Ministry of Agriculture)

Orchards;
122

Field Crops; 260

As stated above, Israeli agriculture is subjected to unique and extreme circum-


stances. Israeli agriculture is characterized by modern and highly mechanized
agriculture with intensive use of irrigation. These all, coupled with the country
being highly populated, result in high levels of stress on the cultivated soils, leading
to intensive soil degradation. More specifically, Israel’s soil degradation arises
primarily from extensive agricultural activities such as frequent tilling, loss of
organic carbon, irrigation with variable water quality, the impact of climate change
on rain patterns, and soil loss due to urbanization and infrastructure. The following
discussion presents Israel’s main drivers for soil degradation [3].

2 Soil Erosion

According to estimates made by the Israeli Ministry of Agriculture and Rural


Development, about 70% of the agricultural land in Israel is in danger of soil erosion
[3]. Of this 60%, almost half are at severe risk for erosion, and about half are
moderate. Studies conducted in Israel have revealed that in areas with medium to
severe potential for erosion, the average annual soil loss is about 4 mm per year [6],
similar to average rates worldwide [7]. This level is higher than the average soil loss
measured in European agricultural areas and an order of magnitude higher than the
upper threshold set by the US Department of Agriculture for Sustainable agriculture.
In regions susceptible to soil erosion in Israel, soil loss is two orders of magnitude
larger than the natural soil formation rate of about 0.02–0.04 mm per year
[7, 8]. Therefore, it is evident that the soil resources in Israel are at high risk.
It is important to emphasize that soil erosion in not always noticed. In many cases,
tillage efforts mask sheet and rill marks in the field [6, 9]. It might take several years
to notice signs of soil loss. The result of destruction of soil structure, loss of organic
262 G. Eshel et al.

matter and nutrients might only be seen when it is too late, i.e. only when the profits
from the agricultural activities are harmed. It is for these reasons that soil degradation
by erosion was not always perceived by the farming community, stakeholders, and
policymakers as a real and immediate threat to soil resources and food security.
Repeated loss of soil top layer, which is the layer most fertile and essential to crop
production, leads to continuous soil degradation. Additionally, soil erosion and loss
of organic matter and nutrients increase greenhouse gas emissions, and reduce
biodiversity [6, 9–12].

2.1 Soil Erosion by Water

Water erosion begins when a water drop strikes the bare soil. Soil particles are then
detached from the surface and subsequently transported down-slope by raindrop
splash or by runoff (overland flow). If the runoff is in thin sheets, sheet flow and
sheet erosion (or interrill) occur. Ellison [13] suggested that for water velocity above
0.3 m s-1 the flow becomes turbulent and causes the formation of soil rills. These
processes occur by water drop impact originating from either rain or overhead
irrigation.
The sensitivity of soil to water erosion is enhanced by leaving the surface of
cultivated soils bare. The common practice of intense tillage results in a break-down
of the natural structure of the upper soil layer. Tillage is practiced in order to prepare
seedbeds, weed control, sanitation, and application of organic manure and fertilizers
[6, 14]. The more intensive the cultivation practiced, the more severe the expected
destruction of the soil structure and soil aggregate stability.

2.2 Soil Erosion by Wind

Soil erosion by wind (i.e., wind erosion) and dust emissions to the atmosphere
significantly impact human environments. These emissions are associated with
agro-environmental hazards, which severely impact crops production caused by
heavy damages to young crops during the planting season due to the mechanical
impact of saltating sand particles and stable soil aggregates. Furthermore, wind
erosion causes considerable nutrient loss and accelerates soil degradation processes
due to cultivation activities that significantly increase the erodible fraction to wind
erosion. These cultivation actions stimulate soil erosivity, affected mainly by
mechanical abrasion of soil aggregates [15].
The most susceptible regions to wind erosion processes are the western Negev
areas where the soils types vary from sandy soil along with the coastal areas and
loess soils abundant along the arid transition zones between the coastal plains and
deserts in the south [16] and the peat soils of the Hula Valley, located in the upper
Jordan catchment. When the land is bare during transition seasons, these regions are
Degradation of Agricultural Lands in Israel 263

affected by accelerated wind shear stress over the soil surface. The shear stress is
defined by the friction velocity (u*), which expresses the boundary layer’s wind
velocity gradient. The threshold wind velocity (u*t), the minimum friction velocity
required for initiating the movement of soil particles, is a crucial parameter control-
ling the frequency and intensity of wind erosion events. Determination of u*t is
complex since this property is affected by several surface factors, such as soil
texture, soil moisture, surface crusts, vegetation cover, or other roughness
elements [17].
In Israel, most agricultural lands are bare during autumn (Sept.–Nov.), when u*
may exceed u*t. Thus, the fields are prone to severe wind erosion events without
being adequately protected by soil conservation measures. In the western Negev
region during the autumn, the dominant wind direction is southwest and its average
maximal wind gusts exceed 9.0 m s-1. On the other hand, in the Hula Valley region,
the dominant wind direction varies from east to east-north-east, with average
maximal wind gusts above 12.0 m s-1. Considering that the average threshold
wind velocity required for soil aggregate detachment is ~6.5 m s-1, large agricultural
fields are exposed to wind shear stress above u*t (Argaman A., Unpublished data).
These dust emissions are associated with agro-environmental hazards that
severely affect crop production through damaging young crops during the seed
emerging season due to the mechanical impact of saltating sand particles and stable
soil aggregates. Furthermore, wind erosion causes considerable nutrient loss and
accelerates soil degradation processes following cultivation activities that signifi-
cantly increase the erodible fraction of soil aggregates to wind erosion.

3 Loss of Organic Carbon

Extensive changes in land use and conversion of natural ecosystems including


forests, prairies, and swamps, in favor of agriculture land occurred in the last few
generations. These led to the oxidation of organic matter and the loss of about
25–75% of soil organic carbon (SOC) that has been emitted into the atmosphere
as carbon dioxide [18]. The SOC is of high importance in the complex soil food web.
SOC is a positive driver of soil health, resilience to climate change, soil structure,
permeability and water retention, and regulating soil temperatures [19]. Soil is one of
the largest reservoirs of world carbon. Most of the SOC is contained in the top 30 cm
of the soil profile [20]. The average time required for the formation of one inch of
soil is 250–500 years. Intensive agriculture accelerates the loss of SOC, damages soil
structure, accelerates soil erosion by water and wind, and violates biodiversity. In
turn, these cause continuous damage to soil fertility and food security. Estimates
show that by 2050, an increase of 70% (or more) in food demand is expected. This
increase in food demand will further increase the stress on soil resources and can
cause intensive soil degradation. The adoption of conservation agriculture
(CA) practices can mitigate soil degradation by improving soil health [10, 21].
264 G. Eshel et al.

A preliminary study was conducted at The Soil Conservation and Drainage


division of the Israeli Ministry of Agriculture in order to estimate the SOC stored
in agricultural land in Israel. The potential for additional carbon storage retention in
the case of CA practices adoption was studied [22]. The estimate was based on SOC
and bulk density data of the top 30 cm of different soil profiles in agricultural fields
[9, 23]. This preliminary study suggests that by shifting all conventional tillage in
Israeli agricultural field to the adoption of CA principles, an amount of six million
tons of carbon can be added to the top 30 cm of the soil over the coming years, thus
reducing atmospheric greenhousehouse gas. This estimate depends on the rate of CA
adoption.

4 Agronomic Approaches for Soil Degradation Mitigation

The best practice for preventing soil degradation by erosion is the adoption of CA
practices. CA practices are based on three pillars: (i) minimal soil disturbance (i.e.,
no or minimal tillage), (ii) permanent organic soil cover, and (iii) diversification of
plant species by crop rotation coupled with a combination of mixed service crops
(cover crops) [24]. However, maintaining permanent soil cover is not commonly
practiced in field crops in Israel. This is because of the high demand for bio-
mass resulting from the ban on the import of coarse food for livestock feed. Thus,
there is no incentive for farmers to leave biomass residue in cultivated fields.
As a result, the widely adopted agronomic measure for soil conservation and
reclamation by the Israeli farmers in rain-fed field crops is the approach of conser-
vation tillage (CT). This approach includes Reduced Tillage for reducing soil
disturbance and manure addition (10 m3 ha-1 y-1) for reclaiming soil fertility
every third or fourth year of the crop rotation.
A recent survey, carried out in 2019, was conducted to examine the degree of
adoption of CT & CA practices in 150,000 ha of orchards and field crops in Israel
[25]. The survey examined the types of tillage practices used and assigned them with
grades as follows. Practice of “Conventional Tillage” was graded as “none,” “reduce
tillage” as “low,” “Minimum Tillage” as “Moderate,” CT as “Good” and CA as
“Best Practice.” This survey showed that for field crops (consisting mostly of rain-
fed crops with crop rotation) in Israel, an area of 11,500 ha (~11% of the surveyed
area) received a grade of “Good” (Fig. 2).
As opposed to field crops, in orchards and vineyards, more and more farmers are
adopting CA practices by eliminating soil disturbance and maintaining permanent
organic soil cover. The common approach for achieving permanent organic soil
cover is by introducing and managing service crops (cover crops). The service crops
can be established by seeding commercial crops (mainly oat) or managing sponta-
neous weeds as a service crop by mowing or rolling. The aforementioned survey also
showed that 8,600 ha of orchards are under CA (accounting for 25% of the surveyed
Orchards) received a grade of “Good” and above [25].
Degradation of Agricultural Lands in Israel 265

Fig. 2 Percentage area of the different groups representing the degree of adoption of Conservation
Agriculture practices in Israel. The data from a survey taken in 2019. Field crops (a), orchards (b)
(Adopted from [25])

The Soil Conservation and Drainage division of the Israeli Ministry of Agricul-
ture and Rural Development supports farmers in several ways to cope with soil
degradation. The support includes tutorials and guidance combined with direct
financial support. In the last 5 years, an average of 4.5 million USD per year have
been allocated for changing agricultural practices, and field planning for applying
physical soil conservation measures combining different elements such as weeding
waterways and channels, terraces and contour benches, buffer strips, and installation
of tile drainage systems in the case of poor drainage and salinity (Maor A. Personal
communication). In addition, the Conservation and Drainage division supports
knowledge development, whether by the Soil Erosion Research Station R&D unit,
supporting and partnering in the Model Farm for Sustainable Agriculture at Neve
Ya’ar, and supporting research through grants from the Chief Scientist of the
Ministry of Agriculture and Rural Development.
Yet, degree of adoption of CA practices in field crops, both irrigated and rain-fed,
and vegetables (Fig. 2) is much lower than that in orchards. The data presented in
Figs. 1 and 2 clearly show that the cultivated soils in Israel are exposed to a
continuous course of degradation. Unless immediate measures such as CA are
employed on a large scale, a tangible danger to food production capacity can be
foreseen.
Among the common soil conservation practices that can be effectively utilized to
mitigate wind erosion, the use of polyacrylamide (PAM) as a soil stabilizing agent
proved highly efficient when applied following seedling [26]. Similar to the case of
soil erosion by water, an effective conservation measure is an application of service
crops (also referred to as cover crops) that increase the zero displacements height and
266 G. Eshel et al.

dramatically diminish the wind surface shear stress over the soil surface, below the
threshold detachment force required for saltation and suspension soil particles.
Albeit the high efficiency of the aforementioned practices in reducing soil
susceptibility to wind erosion, their utilization in agricultural fields is limited. That
is so, mainly due to the questionable profitability of their application over rainfed
areas where water availability (and costs) are a significant limiting factor.

5 Salinization and Sodification by Irrigation

In former years, most of the irrigation water in Israel comprised of freshwater. Use of
saline-sodic water was restricted to localities where shallow aquifers of such mar-
ginal water exist and could easily be pumped (e.g., North Western Negev in the
south of Israel). Irrigation with such water led to severe soil salinization and
sodification [27]. As of the early 1990s, TWW (being more saline and sodic than
their freshwater of origin) has become a significant source of irrigation water [28],
currently reaching ~45% of the water consumed by agriculture, with plans of
reaching ~70% in 2050. Widespread use of TWW exerts critical salinity and sodicity
hazards to the agricultural land in Israel. Long-term TWW irrigation has been
associated with numerous undesired phenomena such as (1) increasing soil sodicity,
especially that of soil sub-surface layers [29]; (2) increasing salinity of both the
saturated and unsaturated zones of the flow domain of the upper 1 m [30]; (3) initiate
hydrophobicity at the soil surface [31]; (4) enhance surface seal formation [28]. In
turn, the aforementioned may lead to degradation of soil structure and functioning,
e.g., weaker aggregates [32], poorer hydraulic conductivity [33], and lower infiltra-
tion rate [34]. Subsequently, the above results in deterioration of the ecosystem
services the soil can provide. In addition, because TWW may contain pharmaceu-
ticals and personal care products, its irrigation could lead to their accumulation in
and contamination of soils and surface and groundwater bodies [35].
In some areas in Israel (mainly on Vertisols), extensive irrigation with low water
quality leads to shallow anthropogenic saline groundwater that enhances the salini-
zation and sodification processes by evaporating the rising capillary waters. These
anthropogenic aquifers needed to be drained to reclaim the salinity and sodicity-
affected fields. Common means for coping with soil salinization in poorly drained
soils and soils with high water tables is by installation of tile drainage system
[36]. These systems consist of horizontal perforated pipes inserted into the soil
(Fig. 3). The systems then produces gravitational drainage paths to high water tables,
draining excess water and salts and from the soil profile to adjacent channels.
Such systems have been found effective in decreasing salinity and sodium
adsorption ratio (SAR) levels and draining excess water. A significant decrease in
salinity is obtained generally after 2 or 3 years from installation [37]. Yet, tile
drainage systems do have some disadvantages. Planning and installing tile drainage
is expensive. Following requests by farmers from the Ministry of Agricultural and
Rural Development for support, more than nine million USD have been invested in
Degradation of Agricultural Lands in Israel 267

Fig. 3 A schematic diagram of lowering water table using tile drainage

various tile drainage systems in cultivated lands in Israel in the last 5 years. These
high costs are met through government support of 50–75%. In addition to the high
costs, the installation procedure involves soil disturbance resulting from intensive
earthwork (Maor A., personal communication). Moreover, the drain outlets have the
potential of polluting the waterways with saline water, fertilizers, and pesticides.
Estimates relying on measurements from two tile drainage system in the Jezreel
Valley showed annual leachate including a total of 184 and 257 ton TDS and 0.82
and 1.8 ton NO3. Tile drainage systems differ greatly in area and efficiency, thus
extrapolating these data to other systems is problematic (Volk E., unpublished data).
Unfortunately, other agronomic approaches such as precise irrigation and fertil-
ization have still not been fully adopted by farmers and policymakers. The advantage
of these solutions is that they are a primary mean for handling the source of the
problem rather than dealing with the symptoms of soil salinization the way tile
drainage does.
A GIS analysis has been performed in order to estimate the extent of risk for soil
salinization in agricultural land as a result of a high water table. Two GIS layers were
used, a layer of Israeli soil associations and one of the agricultural plots. Twenty-two
soil associations from a total of 89 soil associations [38] were selected and catego-
rized by three groups of severity as predictors of salinity and drainage problems
(Table 1). This rationale was based mainly on soil texture and field slope (example:
clayey soils in valleys). This type of division categorized the risk for salinization of
agricultural plots in Israel by its severity.
268 G. Eshel et al.

Table 1 Selected Israeli soil associations [38], categorized by expected severity of salinity
potential
Minor problems
Severe problems expected Moderate problems expected expected
E1 – Hamric alluvial soil and gley H1 – Alluvial Brown A7 – Brown Grumosol
Grumosol
H4 – Hydromorphic Grumosol and H2 – Accumulative Brown B6 – Brown Grumosol
Grumosolic gley and Reddish Brown and Brown Rendezina
Grumosol
H5 – Alluvial Brown Grumosols and H3 – Calcareous accumula- D3 – Basaltic very dark
Hydromorphic Grumosol tive Brown Grumosol and Brown Grumosol
residual dark Brown
H8 – Organic Soils H6 – Light colored chalky D4 – Basaltic reddish
colluvial alluvial soils Brown Grumosol
L2 – Calcareous sierozem H7 – colluvial alluvial soils H9 – Very dark Brown
and Grumosols Bazaltic Grumosol and
Brown Grumosol
L3 – Calcareous sierozem, marly H12 – Gravelly Nazzaz and H10 – Very dark Brown
hydromorphic Grumosol, and highly Brown Hamra Bazaltic Grumosol
calcareous Grumosol
L4 – Lacustrine gley H11 – Reddish Brown
Bazaltic Grumosol
Y7 – Marly Solonchak
Z2 – Solonchak

According to this analysis, 9% of the agricultural land in Israel is expected to


suffer from severe risk for salinization as a result of high water table, and an
additional 33% of the agricultural land is of a moderate risk for salinization as a
result of natural high water tables. In order to validate this analysis, a GIS layer of
existing installed tile drainage systems has been added. The GIS classification shows
an agreement between the plots classified with expected severe problems to plots
containing massive tile drainage systems. Similarly, the plots classified as expected
to have moderate problems contain less massive tile drainage systems, and the plots
classified to have minor expected problems contain small local systems (Fig. 4). This
kind of analysis may serve as a policy tool for future water quality allocation for
irrigation and funds allocation for salinity reclamation.

6 Agriculture Land Loss to Construction


and Infrastructure

Transformation of land use from agricultural land to housing and infrastructure,


mainly transportation and renewable energy, results from the growing population
and elevation of the population welfare in Israel. This is an additional threat to
permanent soil degradation. Furthermore, sealing the soil with infrastructure causes
Degradation of Agricultural Lands in Israel 269

Fig. 4 A map of the severity for salinization risk in Harod and Beit Shean valleys, Israel, and the
layer of installed tile drainage systems (in gray)

a threat to food security and the national water cycle. The pressure for land-use
changes has increased even more dramatically since 2016, when the Israeli govern-
ment approved large-scale housing strategies and established the strategic housing
plan (SHP). A recent study shows that from 2016 to 2040, about 40,000 ha
(an average of 1,660 ha year-1) of agricultural land will be converted for housing
and local infrastructure including residential parks and roads [39]. These numbers
are in agreement with estimates of the Israeli Ministry of Agriculture and rural
development which projected that 1,000 ha of agricultural land had been lost to
urbanization and infrastructure every year in the last decade.

7 Concluding Remarks

Israel is characterized by long hot and dry summers and frequent cycles of drought
years. That, coupled with a shortage of good quality water that forces the use of
marginal water for irrigation, are acute challenges that the Israeli farming community
is facing. Moreover, albeit these harsh conditions, Israeli growers are expected to
support the food security to a rapidly growing population from less than a million at
270 G. Eshel et al.

the initiation of the state to ~nine million nowadays and to a projected doubling of
that in 2050.
Israeli agriculture is modern and intensive, with about half the cultivated area
being irrigated. Consequently, the agricultural soils of Israel are under constant
pressure: on the one hand, land-use changes to construction and infrastructure and,
on the other hand, rapid soil degradation that manifests itself in greater (i) water and
wind erosion, (ii) soil organic matter depletion, and (iii) soil salinization and
sodification.
The importance of protecting agricultural land by minimizing loss for construc-
tion and infrastructure and by employing CA practices combined with improving
soil health has gained increased awareness in recent years by both farmers and
policymakers in order to prevent further soil degradation and build resilience to
the ongoing climate changes, thereby boosting local fresh food production. Yet,
current efforts and actions need to be further encouraged and strengthened to ensure
sustainable use of Israel’s soils so that the needs of future generations can be met.

References

1. Dan J (1983) Soil chronosequences in Israel. Catena 10:287–319. https://doi.org/10.1016/0341-


8162(83)90001-2
2. Yaalon DH (1997) Soils in the Mediterranean region: what makes them different? Catena 28:
157–169. https://doi.org/10.1016/S0341-8162(96)00035-5
3. Itkin D, Bar-Tal A, Crouvi O, Dor M, Erel R, Gross A et al (2022) Soil priorities in Israel.
Geoderma Reg 29:e00505. https://doi.org/10.1016/j.geodrs.2022.e00505
4. Water Authority of Israel (2012) Israel Water Sector Master Plan 2050
5. Central Bureau of Statistic (2020) Israel in figures, selected data from statistical abstract of Israel
6. Eshel G, Egozi R, Goldwasser Y, Kashti Y, Fine P, Hayut E et al (2015) Benefits of growing
potatoes under cover crops in a Mediterranean climate. Agric Ecosyst Environ 211:1–9
7. Montgomery DR (2007) Dirt: the erosion of civilizations. University of California Press,
Berkeley
8. Montanarella L, Pennock DJ, McKenzie N, Badraoui M, Chude V, Baptista I et al (2016)
World’s soils are under threat. Soil 2:79–82. https://doi.org/10.5194/soil-2-79-2016
9. Eshel G, Lifschitz D, Bonfil DJ, Sternberg M (2014) Carbon exchange in rainfed wheat fields:
effects of long-term tillage and fertilization under arid conditions. Agric Ecosyst Environ 195:
112–119. https://doi.org/10.1016/j.agee.2014.05.007
10. Prince S, Von Maltitz G, Zhang F, Byrne K, Driscoll C, Eshel G, et al (2018) Chapter 4: Status
and trends of land degradation and restoration and associated changes in biodiversity and
ecosystem functions, pp 221–338. doi:https://doi.org/10.5281/zenodo.3237392
11. Eshel G, Unc A, Egozi R, Shakartchy E, Doniger T, Steinberger Y et al (2021) Orchard floor
management effect on soil free-living nematode communities. Soil Res. https://doi.org/10.1071/
SR21196
12. Goldshleger N, Ben-Dor E, Lugassi R, Eshel G (2010) Soil degradation monitoring by remote
sensing: examples with three degradation processes. Soil Sci Soc Am J 74:1433–1445. https://
doi.org/10.2136/sssaj2009.0351
13. Ellison WD (1945) Some effects of raindrops and surface flow on soil erosion and infiltration.
Trans Am Geophys Union 26:415–429
14. Eshel G, Fine P, Singer MJ (2007) Total soil carbon and water quality: an implication for carbon
sequestration. Soil Sci Soc Am J 71:397. https://doi.org/10.2136/sssaj2006.0061
Degradation of Agricultural Lands in Israel 271

15. Hagen L (1991) Wind erosion mechanics: abrasion of aggregated soil. https://doi.org/10.13031/
2013.31737
16. Singer A (ed) (2007) Soil classification in Israel. Soils Isr. Springer, Berlin, pp 249–262. https://
doi.org/10.1007/978-3-540-71734-8_10
17. Argaman E, Singer A, Tsoar H (2006) Erodibility of some crust forming soils/sediments from
the Southern Aral Sea Basin as determined in a wind tunnel. Earth Surf Process Landf 31:47–
63. https://doi.org/10.1002/esp.1230
18. Lal R (2011) Sequestering carbon in soils of agro-ecosystems. Food Policy 36:S33–S39. https://
doi.org/10.1016/j.foodpol.2010.12.001
19. Page KL, Dang YP, Dalal RC (2020) The ability of conservation agriculture to conserve soil
organic carbon and the subsequent impact on soil physical, chemical, and biological properties
and yield. Front Sust Food Syst 4:31
20. Lefèvre C, Rekik F, Alcantara V, Wiese L (2017) Soil organic carbon: the hidden
potential. FAO
21. Gomiero T (2016) Soil degradation, land scarcity and food security: reviewing a complex
challenge. Sustainability 8:281. https://doi.org/10.3390/su8030281
22. Shklyar B, Maor A, Etinger E, Ein-Mor E, Eshel G (2021) First estimations of soil organic
carbon sequestration potential in agricultural land of Israel
23. Rinot O, Borisover M, Levy GJ, Eshel G (2021) Fluorescence spectroscopy: a sensitive tool for
identifying land-use and climatic region effects on the characteristics of water-extractable soil
organic matter. Ecol Indic 121:107103. https://doi.org/10.1016/j.ecolind.2020.107103
24. FAO (2017) Conservation agriculture – revised version. FAO, Rome
25. Maor A, Yaacobi B, Ein-Mor E (2020) First Survey of Conservation Tillage (CT) & Conser-
vation Agriculture (CA) adoption in Israel
26. Genis A, Vulfson L, Ben-Asher J (2013) Combating wind erosion of sandy soils and crop
damage in the coastal deserts: wind tunnel experiments. Aeolian Res 9:69–73. https://doi.org/
10.1016/j.aeolia.2012.08.006
27. Hopmans JW, Qureshi AS, Kisekka I, Munns R, Grattan SR, Rengasamy P et al (2021)
Chapter 1 – critical knowledge gaps and research priorities in global soil salinity. In: Sparks
DL (ed) Advances in agronomy, vol 169. Academic, pp 1–191. https://doi.org/10.1016/bs.
agron.2021.03.001
28. Levy GJ, Fine P, Bar-Tal A (eds) (2011) Treated wastewater in agriculture: use and impacts on
the soil, environment and crops.1st edn. Wiley-Blackwell
29. Levy GJ, Fine P, Goldstein D, Azenkot A, Zilberman A, Chazan A et al (2014) Long term
irrigation with treated wastewater (TWW) and soil sodification. Biosyst Eng 128:4–10. https://
doi.org/10.1016/j.biosystemseng.2014.05.004
30. Russo D, Laufer A, Bardhan G, Levy GJ (2015) Salinity control in a clay soil beneath an
orchard irrigated with treated waste water in the presence of a high water table: a numerical
study. J Hydrol 531:198–213. https://doi.org/10.1016/j.jhydrol.2015.04.013
31. Wallach R, Ben-Arie O, Graber ER (2005) Soil water repellency induced by long-term
irrigation with treated sewage effluent. J Environ Qual 34:1910–1920. https://doi.org/10.
2134/jeq2005.0073
32. Paudel I, Bar-Tal A, Levy GJ, Rotbart N, Ephrath JE, Cohen S (2018) Treated wastewater
irrigation: soil variables and grapefruit tree performance. Agric Water Manag 204:126–137.
https://doi.org/10.1016/j.agwat.2018.04.006
33. Bardhan G, Russo D, Goldstein D, Levy GJ (2016) Changes in the hydraulic properties of a clay
soil under long-term irrigation with treated wastewater. Geoderma 264:1–9. https://doi.org/10.
1016/j.geoderma.2015.10.004
34. Assouline S, Mualem Y (1997) Modeling the dynamics of seal formation and its effect on
infiltration as related to soil and rainfall characteristics. Water Resour Res 33:1527. https://doi.
org/10.1029/96WR02674
272 G. Eshel et al.

35. Chefetz B, Mualem T, Ben-Ari J (2008) Sorption and mobility of pharmaceutical compounds in
soil irrigated with reclaimed wastewater. Chemosphere 73:1335–1343. https://doi.org/10.1016/
j.chemosphere.2008.06.070
36. Warrick AW (ed) (2001) Soil physics companion.1st edn. CRC Press
37. Benyamini Y, Mirlas V, Marish S, Gottesman M, Fizik E, Agassi M (2005) A survey of soil
salinity and groundwater level control systems in irrigated fields in the Jezre’el Valley. Israel
Agric Water Manag 76:181–194. https://doi.org/10.1016/j.agwat.2005.01.016
38. Dan Y, Raz Z (1970) The soil Association of Israel
39. Feitelson E, Horowitz-Harel A, Levin N, Mintz Z, Steenekamp G, Zaban S (2021) Haste makes
waste: on the implications of rapid planning in Israel. Land Use Policy 103:105312. https://doi.
org/10.1016/j.landusepol.2021.105312
Agricultural Land Degradation in Kenya

Kevin Z. Mganga

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
2 Land Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
3 Causes of Land Degradation in Arid and Semi-arid Drylands in Kenya . . . . . . . . . . . . . . . . . . 279
3.1 Livestock Overgrazing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
3.2 Introduction of Exotic Invasive Plant Species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
3.3 Soil Nutrient Mining . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
3.4 Climate Variability and Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
3.5 Irrigation Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
3.6 Artisanal and Small-Scale Mining . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
3.7 Increase in Human Population . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
4 Combating Land Degradation in Kenya: A Historical Perspective . . . . . . . . . . . . . . . . . . . . . . . . 284
4.1 The Swynnerton Plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
4.2 Sessional Paper No.10 African Socialism and Its Application to Planning in Kenya 286
4.3 National Policy for the Sustainable Development of Arid and Semi-arid Lands
of Kenya . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
5 Some Strategies to Combat Degradation in Arid and Semi-arid Drylands in Kenya . . . . . . 287
5.1 Conservation Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
5.2 Dryland Agroforestry Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
5.3 Indigenous Grass Reseeding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
5.4 Rangeland Enclosures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

Abstract Arid and semi-arid drylands cover approximately over 80% of Kenya’s
landmass and are home to almost 30 and 70% of the human and livestock
populations, respectively. Pastoral livestock–crop production systems constitute
the main economic activity and source of livelihood. In Kenya, land degradation

K. Z. Mganga (✉)
Copernicus Institute of Sustainable Development, Utrecht University, Netherlands, Utrecht
Department of Agricultural Sciences, University of Helsinki, Helsinki, Finland
e-mail: [email protected]; kevin.mganga@helsinki.fi

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 273
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 273–300, DOI 10.1007/698_2022_929,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 22 December 2022
274 K. Z. Mganga

has widely been considered as a major environmental challenge threatening the


livelihoods of over 12 million people living in degraded lands. Nearly 80% of
Kenya’s landmass is affected by land degradation. It has been estimated that
approximately 30–40% of the arid and semi-arid lands are rapidly being lost through
degradation and an additional 2% have completely been degraded. The total eco-
nomic value of land degradation in Kenya was at 1.3 billion USD annually between
2001 and 2009. Principal degradation processes in the marginal arid and semi-arid
environments in Kenya include salinization, soil compaction, vegetation degrada-
tion, and soil erosion. Soil loss in these dryland zones mainly occurs when vegeta-
tion cover is removed exposing the soil to the agents of erosion, i.e., water and wind.
Arid and semi-arid environments are particularly vulnerable to degradation because
the soils are characterized by a poor structure exacerbated by scarce vegetation
cover. The main causes of land degradation in these dryland environments in
Kenya include overgrazing, woodland deforestation, poor rainfed and irrigation
agricultural practices, and mineral mining activities. Land degradation process in
these marginal lands is aggravated and accelerated by the effects of climate change
and variability (recurrent droughts and floods) and increased human population.
Increased human population has continued to exert tremendous pressure on the
Kenyan dryland resources. However, it also serves as the main driver and valuable
resource in reversing degradation at farm and landscape scales through soil conser-
vation and sustainable land management practices such as conservation agriculture,
dryland agroforestry, indigenous grass reseeding, and rangeland enclosures. These
practices among pastoral and agropastoral communities in the Kenyan drylands have
generally been effective in combating land degradation and enhancing ecosystem
services. Despite the numerous climatic and socioeconomic challenges in
implementing these practices in dryland environments, they remain a viable option
for enhancing sustainable crop and livestock production in the arid and semi-arid
zones in Kenya. Combating land degradation in the Kenyan drylands can best be
achieved through the continued involvement of the community in the selection and
implementation of appropriate practices to address site- and/or landscape-specific
challenges.

Keywords Biological nitrogen fixation, Ecological restoration, Nature-based


solutions, Policy instruments, Rehabilitation, Soil degradation

1 Introduction

Kenya is located in the tropics at around longitude 34°E–42°E and latitude 5°S–5°N
and can be divided into seven broad agro-climatic zones (ACZ) using a moisture
index based on annual rainfall expressed as a percentage of potential evaporation.
Agro-climatic zones I, II, and III (Table 1, Fig. 1) have an index greater than 50%,
account for about 18% of Kenya’s land area, and have a high potential for cropping.
Agricultural Land Degradation in Kenya 275

Table 1 Characteristics of agro-climate zones and farming systems in Kenya


Average
Average annual
Moisture annual potential
index Climate rainfall evaporation
Zone (%) classification (mm) (mm) Vegetation Farming systems
I >80 Humid 1,100– 1,200– Moist Dairy, sheep, cof-
2,700 2,000 forest fee, tea, maize,
sugarcane
II 65–80 Sub-humid 1,000– 1,300– Moist and Maize, pyrethrum,
1,600 2,100 dry forest wheat, coffee,
sugarcane
III 50–65 Semi-humid 800– 1,450– Dry forest Wheat, maize, bar-
1,400 2,200 and moist ley, coffee, cotton,
woodland coconut, cassava
IV 40–50 Semi-humid 600– 1,550– Dry wood- Ranching, cattle,
to semi-arid 1,000 2,200 land and sheep, barley, sun-
bushland flower, maize, cot-
ton, cashew nuts,
cassava
V 25–40 Semi-arid 450–900 1,650– Bushland Ranching, live-
2,300 stock, sorghum,
millet
VI 15–25 Arid 300–550 1,900– Bushland Ranching,
2,400 and nomadism
scrubland
VII <15 Very arid 150–350 2,100– Desert Nomadism and
2,500 scrub shifting grazing
Source: Sombroek et al. [85]

The semi-humid to arid regions (zones IV, V, VI, and VII) have indexes of less than
50% [1]. Arid and semi-arid lands (ASALs) of Kenya cover over 80% of the
country’s land mass and are home to nearly 30 and 70% of the human and livestock
populations, respectively [2].
Arid and semi-arid drylands in Kenya receive very low and erratic bimodal
rainfall that is highly variable. On average, annual rainfall ranges from 150 to
450 mm and 500 to 850 mm in the arid and semi-arid lands, respectively. However,
these annual rainfall figures can be misleading. This is because rainfall in the ASALs
of Kenya is highly variable in space and time with some years characterized by
above and below average annual rainfall. Subsequently, the inter-annual rainfall in
these environments can vary from 50 to 100% in arid lands and 20–30% in the semi-
arid lands. Abrupt shifts in Intertropical Convergence Zone (ITCZ) and the arrival of
steady deviate meridional winds like Somali jet in January to February, and in June–
September, contribute to the low rainfall experienced in the ASALs of Kenya
[3]. The ASALs are characterized by high ambient temperatures with a wide diurnal
range. Ninety percent of the arid and semi-arid areas lie below 1,260 m, and mean
276 K. Z. Mganga

Fig. 1 Agro-climatic zones in Kenya (Source, Kenya Soil Survey)

annual temperatures range from 22°C to 40°C [1]. In most areas, evapotranspiration
rates are more than twice the annual rainfall.
Dryland soils in Kenya are generally shallow, highly variable, low in fertility, and
subject to compaction, capping, and erosion. Soil nutrient deficiency is one of the
factors that limit cowpea yields in the soils of dryland of eastern Kenya, which are
commonly deficient in nitrogen (N) and phosphorus (P) [4]. Although N and P
contents are often less than optimal, they are considered adequate for plant growth
Agricultural Land Degradation in Kenya 277

and crop production. Few arid and semi-arid lands in Kenya have volcanic soils and
alluvial deposits which are suitable for crop production. For example, in the dry
plains (keew) below the Kerio Valley escarpment the soil parent material is domi-
nated by sandy, silty, or clayey colluvial and alluvial deposits [5]. Moreover, heavy
clays, e.g., Vertisols (black cotton soils), are commonly found in Kenyan drylands.
However, cultivation in such soils remains a challenge because of poor workability
and salinization. In central Turkana, north-western Kenya, soil salinity affects
15–20% of the arid rangelands [6].
Natural vegetation in the ASALs of Kenya is diverse, widely varied, and char-
acterized by open savannahs dominated by perennial and annual herbaceous plants
interspersed with woody tree species, dwarf shrubs, and bush species. Some of the
common woody species with a wide ecological range include Acacia tortilis that
occurs mainly in riparian zones, A. senegal mainly occurring on hilly and rocky sites,
A. reficiens on non-riparian sites with fine soils, and A. drepanolobium in Vertisol-
dominated drylands. Some of the common native perennial grasses that dominate the
herbaceous vegetation layer and are a key source of forage for livestock and wild
herbivores include the drought-tolerant Themeda triandra, Cenchrus ciliaris,
Enteropogon macrostachyus, Eragrostis superba, and Chloris roxburghiana
[7]. In addition to the indigenous plant species, some introduced plant species are
now widespread, modifying the vegetation structure in arid and semi-arid landscapes
in Kenya. For example, the exotic Prosopis juliflora introduced in the Kenyan
drylands to combat desertification has increasingly become an important invader
species especially in the riverine floodplains, along roads and around human
settlement [8].
Generally, four (4) broad land use patterns characterize Kenyan drylands. Pasto-
ralism is the key land use system and dominates in the arid lands. Subsistence
pastoralism puts more emphasis on milk and milk-based diets than meat production.
Pastoralism in the arid drylands is characterized by large livestock (cattle, goat,
sheep, and camel) herds with relative take-off. Traditional nomadic pastoralism has
been found to be less risky than dryland agriculture, more economical and less
harmful to the environment [9]. Rainfed agriculture in these environments is scarce
and erratically practiced, though irrigation may be locally important (ecozones V,
VI, and VII). Pastoral-agricultural interface where pastoral lands have a potential for
cultivation is practiced in ecozones IV, V, and VI, such as the Maasai rangelands.
Commercial livestock ranches are present in a number of land use types, though
most are in the agropastoral and marginal ecozones IV, V, and VI agricultural lands.
Dryland agriculture is found in many ecozones in the Kenya drylands and is
characterized by crop–livestock production, with livestock as an important compo-
nent of land use. The most common crops include different varieties of drought-
tolerant cereals and pulses such as maize, sorghum, millet, beans, cowpeas, pigeon
peas, and mung bean [7]. Additionally, wildlife conservation areas, e.g., national
parks, nature reserves, and conservancies, are largely found in the Kenyan drylands.
This chapter gives an overview of land degradation in Kenya, especially in
dryland environments. The main aims are to (a) discuss the various forms and causes
of land degradation in arid and semi-arid lands in Kenya; (b) outline the historical
278 K. Z. Mganga

perspective of combating land degradation in Kenya; and (c) give a synopsis of some
contemporary soil conservation, rehabilitation, and restoration strategies and actions
being practiced to halt and reverse land degradation in the Kenyan drylands.

2 Land Degradation

Land degradation implies reduction of resource potential by one or a combination of


processes acting on the land. These processes include water erosion, wind erosion,
and sedimentation by those agents, long-term reduction in the amount or diversity of
natural vegetation, and salinization and sodication. Degradation of natural vegeta-
tion is a worldwide phenomenon [10]. Land degradation is also described as the
aggregate diminution of the productive potential of the land, including its major uses
(rainfed, arable, irrigation, and rangeland), its farming systems (e.g., small-holder
subsistence), and its value as an economic resource. This link between land degra-
dation and its effect on land use is central to nearly all published definitions of land
degradation. Thus, a degraded land can be defined as land which due to natural
processes or human activity is no longer able to effectively sustain an economic
function and/or the original ecological function.
Under the United Nations Convention to Combat Desertification (UNCCD),
desertification is defined as land degradation in arid, semi-arid, and dry sub-humid
areas, resulting from various factors including climate variations and human activ-
ities. Drylands are considered Earth’s largest biome and cover approximately 41%
(~61 mil km2) of the Earth’s surface [11]. Arid and semi-arid lands are defined by
precipitation as areas that experience annual precipitation (P) of 25 mm ≤ P < 250 mm
and 250 mm ≤ P < 500 mm, respectively [12]. These areas are characterized by
short growing periods (1–74 and 75–119 growing days, respectively) and the rainfall
patterns are generally unpredictable and highly variable in space and time. Thus,
these marginal lands present unique challenges for crop production. Moreover,
dryland environments are more susceptible to land degradation because they have
poor soil structure exacerbated by scarce and depleted vegetation cover [13, 14].
The problem of land degradation is difficult to grasp in its totality. Degradation
indicators are commonly used to better understand the land degradation problem.
Indicators are variables which may show that land degradation has taken place and
are not necessarily the actual degradation itself. Some of the indicators of drylands
degradation include poor soil cover, land cover change, e.g., dominance of undesir-
able plant species, low soil quality and erosion of topsoil, and loss of land produc-
tivity [15]. Indicators of land degradation can generally be grouped into three broad
categories, i.e., biophysical indicators (degradation of soil, water, and vegetation
cover) [16], socioeconomic indicators (poverty and food insecurity) [17], and
institutional indicators (failures in the public/government, private/market, civil/com-
munity sectors) [18]. Land degradation is normally evident in a decline in produc-
tivity, a loss of biodiversity, and an increasing rate of erosion [10]. It is also possible
to conclude that dryland degradation from excellent to poor under arid and semi-arid
Agricultural Land Degradation in Kenya 279

conditions results in a subsequent but proportional increase in dominance of unpal-


atable, undesirable (increasers), and invader species over the desirable (decreasers)
vegetation types [19]. However, the condition of the soil is one of the best indicators
of land degradation since it integrates a variety of important processes, particularly
vegetation growth, overland flow of water, infiltration, land use, and management
practices.

3 Causes of Land Degradation in Arid and Semi-arid


Drylands in Kenya

In Kenya, land degradation is pronounced mainly in the Eastern and North Eastern
regions [20]. This can probably be due to low and variable rainfall, depleted
vegetation cover during the dry seasons, bare soils, and nature of the terrain that
increases vulnerability to degradation (Fig. 2). The extent of land degradation in the
country is shown in Fig. 3. Land degradation in the arid and semi-arid drylands of
Kenya has been attributed to both climatic and anthropogenic factors, yet it is still
unclear which factor is the main driver of the degradation process. Some researchers
consider climate to be the major contributor to the degradation processes, with
human factors playing a relatively minor supporting role. Other researchers reverse
the significance of these two factors and associate land degradation and erosion in
arid and semi-lands with human activities. A third category of researchers blame
climate and humans more or less equally [13]. The nature of this uncertainty
suggests that the processes that lead to land degradation involve complex interac-
tions between societal factors, e.g., poor land management and increasing population
pressures, and natural climatic factors, e.g., cyclical and short-term droughts [21].

3.1 Livestock Overgrazing

Overgrazing by livestock remains a major contributor of environmental change in


arid and semi-arid drylands in Kenya. Despite the complex nature of relationships
between pastoralists, livestock, and the African dryland environment, dominant and
popular narratives have portrayed traditional nomadic pastoralism as the main cause
of “overgrazing” and “environmental degradation” in arid and semi-arid drylands in
Africa [22]. However, this generalization is misleading and often inaccurate. This is
because land reducing mechanisms, such as evictions from grazing lands, agricul-
tural expansion to marginal drylands, and sedentarization, have curtailed traditional
adaptive strategies of pastoralists, e.g., mobility [23]. This has led to a decline in
resource availability and limited the spaces for movement and contributed to severe
degradation. Sedentarization of pastoralists in arid and semi-arid lands in Kenya is
commonly concentrated around watering points. Radial gradients of grazing
280 K. Z. Mganga

N
W E
S

LEGEND
very severe hazard
severe hazard
moderate hazard
slight hazard
very slight hazard
non-hazard
Water body

70 0 70 140 Kilometers

Fig. 2 Land degradation hazard areas in Kenya (Source: Kenya Soil Survey)

intensities around water points create patches of severely degraded soils and vege-
tation that contribute to reduced plant cover, increased exposure of bare soil and
runoff, reduction in rainwater use efficiency, and loss of productivity and diversity.
For example, the degraded semi-arid rangelands of Lake Baringo Basin in Kenya
inhabited by Il Chamus pastoralists are characterized by a lack of vegetation cover,
Agricultural Land Degradation in Kenya 281

Fig. 3 Typical dryland environment in Kenya characterized by depleted vegetation cover and bare
soils (Photo credit: Kevin Z. Mganga)

increased soil erosion, invasion of alien plant species, and bush encroachment,
mainly by Prosopis juliflora [24].

3.2 Introduction of Exotic Invasive Plant Species

Prosopis juliflora was introduced in the Kenyan drylands to combat desertification


and improve livelihood as a source of timber and fuel. However, because of its
invasive nature, pastoralist and agropastoralist communities in the Kenyan ASALs
are producing charcoal from P. juliflora for income generation, diversification of
livelihood as an adaptation to the impacts of climate change, characterized by crop
failures and livestock mortality, and saving pastures for their livestock. Prosopis
juliflora is one the major species for charcoal production because its wood has a very
good heat of combustion due to its high carbon and lignin contents, yielding charcoal
of a high calorific value ranging between 24 and 33 MJ kg-1 [25]. Likewise, in
addition to the exotic and invasive species like P. juliflora, tree species native to the
Kenyan drylands are also used for charcoal production. In Kenya, charcoal is mostly
produced from hardwood indigenous species such as Acacia tortilis, A. nilotica, A
senegal, A. mellifera, A. polyacantha, and A. xanthophloea, which are native to the
282 K. Z. Mganga

drylands. They are the most widely used and preferred species because of their high-
density heartwood and calorific value. For example, A. nilotica, A. tortilis and
A. xanthophloea have estimated calorific values of 20 MJ kg-1 [26],
19–20 MJ kg-1 [27], and 33 MJ kg-1 [28], respectively. Subsequently, sourcing
charcoal from these dryland zones, which are a major source of charcoal in Kenya,
has contributed to the destruction of the native vegetation leading to severe land
degradation.

3.3 Soil Nutrient Mining

Sedentarization of agropastoral and pastoral communities in the ASALs in Kenya


has also partly contributed to increased crop cultivation, an adaptation strategy
against climate variability and change. Low input form of crop production in these
marginal drylands characterized by fragile soils, weak soil structure, and scarce
vegetation cover, especially during fallow periods, has contributed to accelerated
soil degradation [29]. “Soil nutrient mining” in croplands is an important driver of
degradation in agricultural landscapes in ASALs. Croplands with poor soil fertility
and land management practices in Kenyan drylands tend to suffer the most from soil
nutrient depletion, because farmers lack the capacity to sufficiently replenish nutrient
losses through incorporation of crop residues, livestock manures, and inorganic
fertilizers.
Past studies have demonstrated that nutrient mining is occurring for nitrogen,
phosphorus, and potassium in arid and semi-arid croplands in Kenya [30, 31]. Crop
production in these marginal zones is mostly based on mining of inherent fertility.
Replacing mined nutrients in a typical semi-arid environment in Kenya would cost
about US$110–121 for one hectare of land annually [30]. Often, pastoral and
agropastoralists in these zones do not have the financial capacity to cater for these
extra costs. Subsequently, this puts sustainable agricultural production at stake. In
addition to nutrient mining, inappropriate land management and scarce vegetation
cover combined with fragile soils in agricultural lands in the Kenyan drylands have
contributed to massive soil loss through erosion. Annual potential soil erosion from
the River Perkerra Catchment (1,207 km2) in a semi-arid environment in Kenya has
been estimated to be 1.73 million tonnes [32]. Furthermore, applying the WEPP and
EROSION 3D physical models, estimated erosion rates and sediment yields at the
Mara River Basin were significantly higher for cultivated lands (120 t ha-1 year-1)
than in natural bush land (7 t ha-1 year-1) or grasslands (3 t ha-1 year-1) [33]. These
studies demonstrate that poor agricultural practices in arid and semi-arid zones in
Kenya are primarily attributed to multiple socioeconomic factors that limit the
adoption of sustainable land use policies and practices by local scale among pasto-
ralists and agropastoralists.
Agricultural Land Degradation in Kenya 283

3.4 Climate Variability and Change

Land degradation in the Kenyan drylands is also exacerbated and accelerated by


climate variability and change through increased recurrent droughts and floods. The
ASALs in Kenya have faced increasing drought frequency and intensity since the
1960s, making these zones to be among of the most vulnerable and drought-prone
regions in the country [34]. Moreover, these drylands have the highest incidence of
poverty (approximately 65%) in Kenya [35]. Frequent droughts trigger catastrophic
events, e.g., crop failures, famine, and livestock mortality that significantly diminish
adaptive capacity of pastoral and agropastoral communities inhabiting these dryland
environments. Moreover, the changing bimodal rainfall pattern in the ASALs is
often characterized by alternating massive floods after a severe drought season. The
impacts of the flash floods include livestock mortality, crops losses, and massive
erosion. Subsequently, recurrent droughts and flash floods in the Kenyan drylands
compounded with socioeconomic limitations, e.g., poverty, have accelerated the
land degradation process and have led to persistently unstable and declining crop and
livestock productivity, thus a perennial threat to food security [2, 8, 17].

3.5 Irrigation Agriculture

Over the years, the Kenyan government has continually established large-scale
irrigation schemes in the ASALs, e.g., Hola, Perkerra, Bura, Turkana, and more
recently the Galana-Kulalu Food Security Project. Kenya has the third largest
estimated potential for irrigation development (more than 0.5 million ha) in SSA
drylands [36]. Likewise, pastoralists and agropastoralists inhabiting drylands in
Kenya undertake different strategies to manage weather unreliability, thus enhancing
their adaptive capacity to respond to climate variability and change. For example,
since the late 1960s, pastoralists in the dry regions were exposed to and adopted
flood-retreat crop production [37]. These efforts are aimed at addressing the recur-
rent problem of food shortages that cannot be solved through rainfed agricultural
production alone, especially in the drylands, where water scarcity is the main
constraint to agricultural development. Moreover, in these dryland zones, natural
water resources used for irrigation are also predominantly saline [38].
Water logging and irrigation of these drylands have exposed the soils to degra-
dation through salinization [5]. In Kenya, salinization is estimated to affect 30% of
irrigated lands and has contributed to loss of around 27% of irrigated land over the
last century [20, 39]. Salinity is an environmental and economic problem especially
in the drylands as the productivity of soils is gradually reduced by rising saline
groundwater. Furthermore, the use of saline groundwater for irrigation in these
marginal drylands increases the concentration of salts in the rhizosphere, especially
when drainage is restricted and leaching of the salts inadequate. Subsequently, this
forces the rural communities to adjust their enterprise (crop–livestock) mix and/or
invest in mitigation measures at local scale.
284 K. Z. Mganga

3.6 Artisanal and Small-Scale Mining

In addition to small-holder irrigation schemes, pastoral and agropastoral communi-


ties in the Kenyan drylands are also engaging in artisanal and small-scale mining,
especially in the dry season. This is aimed at diversifying their sources of income to
guard against risks associated with rainfed agriculture and livestock production
[40]. Though well intended, mining activities in the arid and semi-arid drylands in
Kenya have often led to degradation mainly due to the poor mining technologies
used by the artisanal miners who dominate the mining of gemstones [41]. Mining
activities in semi-arid Taita-Taveta County have led to massively eroded gullies and
loss of biodiversity due to clearing of native vegetation without any rehabilitation
[42]. Additionally, mine pits and excavated soils render previously productive arid
and semi-arid rangelands unsuitable for grazing, thereby reducing the availability of
the land for livestock production, a key source of livelihood among pastoralists.

3.7 Increase in Human Population

Continuous growth of human population boosts land degradation. Conversion of


arid and semi-arid rangelands into less productive croplands has resulted in the loss
of valuable natural ecosystems. Increasing human population coupled with influx
migration from non-ASALs continues to exert tremendous pressure on the Kenyan
drylands. Although human population growth is listed as top among the drivers of
change in the arid and semi-arid rangelands in Kenya [13, 43], a rapid expansion of
human population on its own is not necessarily related to land degradation. Rather, it
is what the human population does to the land that determines the extent of land
degradation. On the contrary, increased human population can be a major resource in
reversing land degradation. For example, population growth in some ASALs in
Kenya, e.g., West Pokot and Baringo, has resulted in moderate land scarcity and
increased labor availability. This has subsequently triggered the motivation for soil
conservation efforts and sustainable land management [44].

4 Combating Land Degradation in Kenya: A Historical


Perspective

Kenya has a long history of activities to combat desertification and mitigate the
effects of drought dating back to the 1940s and has continually made numerous
attempts to rehabilitate degraded drylands. Such initiatives are aimed at increasing
vegetation cover, enhance biodiversity, and provide ecosystem services [45]. The
Kenyan government aimed at trying to reverse the trend of loss of vegetation cover,
which in turn leads to loss of soil fertility and soil erosion. Soil erosion is the single
most visible and notorious form of environmental degradation in the study area.
Agricultural Land Degradation in Kenya 285

Over the past three to four decades, drylands in Kenya have experienced some
unprecedented land degradation [13].
African drylands, similar to those occurring in the arid and semi-arid zones in
Kenya, are especially susceptible to degradation because they have a poor soil
structure, worsened by scarce vegetation cover [14]. Degradation in African dry-
lands is well exemplified by declining soil fertility, loss of soil biodiversity, and soil
erosion [10]. Land degradation in the arid and semi-arid drylands in Kenya has been
extensively considered a major environmental problem [13, 46, 47].
Despite their susceptibility to degradation, arid and semi-arid drylands are resil-
ient ecosystems and thus responsive to restoration and rehabilitation. Past restoration
and rehabilitation programs undertaken in dryland environments in Kenya, e.g.,
Baringo, Machakos, Makueni, and Kitui counties from the 1930s to present, suggest
that many of the degraded sites, if fenced and protected, are likely to recover rather
quickly and dramatically. However, until such measures are taken, many areas will
produce less fodder, food, and other goods and services essential to rural life. Several
bodies including governmental, inter-governmental, non-governmental, and donor
agencies have been involved in community initiatives to combat land degradation in
the arid and semi-arid drylands in Kenya.
The African Land Development Board (ALDEV) was created in 1946 to address
the serious effects of land degradation. The Kenya Rangeland Ecological Monitoring
Unit (KREMU) was established in 1975 to monitor ecological changes in the
drylands. Most activities to combat land degradation in Kenya have fallen short of
the desired expectations because of among other reasons sectoral approach,
uncoordinated funding, inadequate policies, and inadequate involvement of the
local communities and not taking into account the historical background and cogni-
zance of the lessons learnt in the past attempts to combat land degradation in dryland
environments in Kenya.
The management of Kenya’s drylands has been guided by several cross-sectional
policy instruments most of which came to force long before Kenya ratified the
Convention to Combat Desertification (CCD). Kenya ratified the convention in
1997 with a view to joining the World Community to combat desertification and
land degradation. Some of the policy frameworks and instruments include the
Swynnerton Plan (1954), Sessional Paper No. 10 on African Socialism and its
Application to Planning in Kenya (1965), District Focus for Rural Development
(1982), Sessional Paper No. 1 on Economic Management for Renewed Growth
(1986), Development Policy for Arid and Semi-arid Areas (1992), the National
Environmental Action Plan (1994), Sessional Paper No. 1 (1994) on Recovery and
Sustainable Development to the year 2010, National Development Plans for
1974–1979, 1980–1984, 1985–1989, 1990–1992, 1993–1997, and 1998–2002,
Sessional Paper No. 6 on Environment and Development (1999), Poverty Reduction
Strategy Paper (2001), and National Policy for the Sustainable Development of Arid
and Semi-arid Lands of Kenya (2004). The listed and other policy frameworks and
instruments are available in the Kenya Institute for Public Policy Research and
Analysis Public Policy Repository (KIPPRA PPR) http://repository.kippra.or.ke/
handle/123456789/17 website. Out of all these policy instruments, the following
have had the greatest impact.
286 K. Z. Mganga

4.1 The Swynnerton Plan

During the colonial period, much of Kenya was divided into independent reserves
and high potential areas for colonial settlement agriculture. The persistent low
productivity on the reserves and a growing political insurrection in some parts of
Kenya led to a liberal proposal for land-tenure reform, which continues to shape the
evolving landscape in Ukambani and all of Kenya. The Swynnerton Plan of 1954
was supposed to address African land problems by reforming land tenure, consol-
idating fragmented holdings, issuing freehold title, intensifying and developing
African agriculture, providing access to credit, and removing restrictions on growing
crops for export. It consisted of a three-phase program:
1. Land adjudication to “phase out” customary tenure
2. Land consolidation into one block per household to eliminate small, dispersed
parcels, to allow greater specialization, and to realize economies of scale in cash
crop production
3. Land registration to provide for security of ownership and to establish a land
market
Overall, the aim was to facilitate increased investment and employment in
agriculture and to increase rural incomes and the “productivity” of land [48]. The
plan was predicated on an assumption that explicitly “successful” or wealthy African
farmers would “be able to acquire more land and bad or poor farmers less, creating a
landed and a landless class.”
The Swynnerton Plan emphasized the need to de-stock the pastoralists in the
marginal areas and introduce grazing control. In addition, it also aimed at providing
the African farmer with secure title to private property so as to encourage him invest
his labor and profits into the development of his farm. This minimized the destruc-
tion of the environment, thus combating land degradation. (KIPPRA PPR http://
repository.kippra.or.ke/handle/123456789/2712)

4.2 Sessional Paper No.10 African Socialism and Its


Application to Planning in Kenya

This Sessional Paper was drafted immediately after independence to facilitate the
so-called Africanization of the Kenyan economy and public service. It was aimed at
ensuring a rapid economic development and social progress for all Kenyans. In the
phrase “African Socialism,” the word “African” was not introduced to describe a
continent to which a foreign ideology was to be transplanted. It was meant to convey
the African roots of a system that was in itself African in its characteristics. African
Socialism is a term describing an African political and economic system that is
positively African not being imported from any country or being a blueprint of any
foreign ideology. The principal conditions the system had to satisfy were to draw on
Agricultural Land Degradation in Kenya 287

the best of the African traditions, be adaptable to new and rapidly changing circum-
stances, and not rest for its success on a satellite relationship with any other country
or group of countries. Under African socialism, the power to control resource use
resided with the state. Idle and mismanaged farms owned by the Kenyan citizens or
foreigners were not permitted. Production and yields were also influenced by
government extension officers and research. All this ensured the conservation of
our heritage for the future generations through the adoption and implementation of
policies designed to conserve natural resources. (KIPPRA PPR http://repository.
kippra.or.ke/xmlui/handle/123456789/2345).

4.3 National Policy for the Sustainable Development of Arid


and Semi-arid Lands of Kenya

This policy document was formulated through a participatory and consultative


process, which was spearheaded by Arid Lands Resource Management Project and
United Nations Development Programme (UNDP). The process involved relevant
stakeholders including government departments and research institutions, e.g., ILRI,
KARI, KEFRI, ICRAF, and KEMRI, among others. The objective of the policy is to
provide a coherent and practical framework for the implementation and realization of
a new vision for ASAL development in Kenya. This policy reflects the government’s
commitment to overcome challenges facing the arid and semi-arid drylands and was
aimed at reversing the negative trends prevailing in the ASALs, e.g., land degrada-
tion, in order to uplift the socioeconomic welfare of ASAL inhabitants. According to
the policy document, one of the weaknesses of communal land tenure is that it does
not confer adequate incentives and sanctions for efficient utilization and manage-
ment of common property resources, which leads to what is commonly referred to as
the “tragedy of the commons.” In trying to address this challenge, the policy seeks to
improve ASAL land tenure and land use policies. (KIPPRA PPR http://repository.
kippra.or.ke/handle/123456789/1020)

5 Some Strategies to Combat Degradation in Arid


and Semi-arid Drylands in Kenya

5.1 Conservation Agriculture

Crop production in Kenyan drylands is predominantly based on conventional agri-


cultural practices such as soil tillage and inversion, feeding crop residue to livestock,
and low level of fertilization, leading to soil degradation. Conservation agriculture,
defined as a system of multiple agronomic practices, e.g., reduced tillage or no
tillage, permanent organic soil cover by retaining crop residues, crop rotations, and
288 K. Z. Mganga

intercropping [49], has demonstrated to have great potential for combating land
degradation in crop lands. These agronomic practices result in better soil structure
and have higher SOC compared to conventional tillage practice, especially at the top
soil depths [50]. Moreover, SOM accumulation under no tillage compared to
conventional tillage practices contributes to important improvements in soil quality,
soil fertility, and C sequestration. Therefore, conservation agricultural practices have
a great potential to combat land degradation in dryland cropping systems in Kenya,
characterized by poor soil structure.
A larger percentage of dry aggregates and less wind erodible fraction in no-tillage
practices decrease wind erosion. Erosion by water can also be reduced dramatically
due to the effects of conservation agriculture practices on soil properties and
processes that enhance rainwater infiltration and reduce surface runoff [49]. Higher
macro-aggregates and aggregation indices have been observed in no-till and com-
pared to conventional tillage systems in semi-arid environment in Kenya demon-
strate the potential of conservation agriculture in controlling soil erosion and
enhancing soil water content [51]. Furthermore, no tillage with residue retention
has also contributed to high soil water content and nitrogen use efficiency in a
dryland environment in Kenya similar to other regions in sub-Saharan Africa where
conservation agriculture-based practices have been reported to improve soil water
retention [52, 53].
Intercropping cereals and legumes is a critical component of conservation agri-
culture. This is because it contributes toward reversing land degradation and increas-
ing food production in arid and semi-arid regions in Kenya. Intercropping has been
identified as a sustainable agronomic strategy that offers greater stability than
monocropping in terms of enhancing soil fertility and crop yields [54]. Cajanus
cajan (L.) Millsp. (pigeonpea) has demonstrated a great potential for intercropping
with drought-tolerant varieties of cereal crops such as Zea mays (L.) (maize),
Sorghum bicolor (L.) (sorghum), Pennisetum glaucum (L.) R. Br. (pearl millet),
and Eleusine coracana (L.) Gaertn. (finger millet) in the Kenyan drylands. This is
mainly attributed to its drought tolerance and contribution to household nutrition,
income generation, and enhanced crop productivity [55, 56]. Additionally, C. cajan
is incorporated in intercropping systems because of its slow initial growth, ability to
maintain growth and produce grain much later after harvesting the cereal crop, high
N-fixing capacity, and mineralization of sparingly soluble P on degraded sites
[57]. Three pigeonpea varieties fixed 60–70 kg N ha-1 in a maize–pigeonpea
intercropping system in semi-arid Kenya [58]. In another study in semi-arid
Kenya, pigeonpea–maize intercrop recycled 27 kg N and 1.6 kg P ha-1 annually
[55]. Incorporation of pigeonpea in intercropping systems has the ability to biolog-
ically fix N in small-scale croplands across different soil types and rehabilitate
degraded soils through input of 70–100 kg of N per hectare annually [57]. In another
study, cowpea fixed 50% of their N, and so provided a net input of N into a maize–
cowpea intercrop system in four growing seasons in a dryland zone in Kenya
[59]. These studies demonstrate that intercropping cereals and legumes, e.g., pigeon
peas and cowpeas, could ameliorate N deficiency that is a primary limiting factor for
crop production in the arid and semi-arid zones in Kenya.
Agricultural Land Degradation in Kenya 289

Crop rotations, i.e., the sequential and systematic cultivation of crops in succes-
sion, also play a key role in conservation agriculture. Adoption of crop rotations has
mainly been linked to its contribution to increased crop yield attributed to increased
soil fertility (notably rotations incorporating legumes), enhanced soil structure, and
pest and weed control [60]. Crop rotations with drought-tolerant annual legumes,
e.g., Vigna unguiculata (cowpeas), have the potential to increase soil N concentra-
tion through biological N-fixation in arid and semi-arid zones in Kenya. Previous
studies have demonstrated that inorganic fertilizers or organic manure do not solve
the challenges of soil degradation and nutrient depletion unless when used in
combination with other agronomic practices [61, 62]. Crop rotations of lablab/
pigeon pea or a cowpea/pigeon pea combination with maize or sorghum on an
infertile soil in semi-arid zone in Kenya gave an extra grain production of
300–400 kg ha-1 and the legumes contributed about 40 kg N ha-1 [63]. These
studies demonstrate that conservation agriculture agronomic practices have great
potential to mitigate the impacts of soil nutrient depletion and contribute to combat-
ing land degradation in arid and semi-arid drylands in Kenya.

5.2 Dryland Agroforestry Systems

Deliberate incorporation of drought-tolerant perennial trees and shrub species and


food crops in intimate combination with one another is a strategy used by
agropastoralists to combat land degradation in the Kenyan arid and semi-arid
drylands. Cultivation of trees in dryland agricultural systems improves vegetation
cover in croplands and enhances soil fertility through various processes. These
include biological nitrogen fixation by rhizobial bacteria in root nodules of
e.g. leguminous tree species, nutrient uptake by trees from deeper soil horizons
beyond the crop rooting zones and subsequent surface deposition as litter fall,
capture of nutrients which would otherwise be lost through leaching, improved
soil physical characteristics (e.g., aggregation and enhanced rainwater infiltration
via stem flow) due to increased soil organic matter (SOM) content, reduction in
aluminum (Al) toxicity, and low pH through the facilitation of cycling of bases and
the production of complex metabolic Al compounds and improved soil microbial
activity of soil organisms through a cooler and moist microclimate [64]. Addition-
ally, trees in croplands form physical barriers that minimize runoff, enhance infil-
tration, and trap soil sediments, thus reducing soil erosion and protecting soil
aggregates by reducing the kinetic energy of raindrops.
Traditional dryland agroforestry systems in Kenya are characterized by scattered
trees on croplands, farm boundaries tree plantations, wood lots and block planta-
tions, native trees on arid and semi-arid range lands, and live hedges/fences.
Drought-tolerant multi-purpose trees for soil conservation and reclamation, fruit
production, source of timber and fuel wood, livestock and bee forage, fencing,
shade, and N-fixation are the most preferred species in agrisilvicultural and
agrosilvopastoral systems in the Kenyan ASALs [65]. Contour hedgerows of
290 K. Z. Mganga

Senna siamea and mulch have demonstrated to be the most effective system in
reducing soil loss from just over 100 to only 2 t ha-1 and runoff from just below
100 to 20 mm as compared to a maize and cowpea monocropping system in a semi-
arid environment in southeastern Kenya [66]. Nutrient cycling in an agroforestry
system with runoff irrigation in an arid environment in Kenya demonstrated that the
incorporation of Acacia saligna trees into cropland diminished nutrient leaching
compared to S. bicolor monocropping system, leading to a higher nutrient use
efficiency [67].
Additionally, tree species native to arid and semi-arid drylands in Kenya and
incorporated in dryland agroforestry systems, e.g., Acacia senegal, A. nilotica,
A. seyal, A. tortilis, and Salvadora persica, have displayed sustained growth and
great potential in rehabilitating sites affected by soil salinity in the arid zone of
central Turkana, north-western Kenya [6]. Studies from other drylands regions in
Eastern Africa have also demonstrated the great potential of agroforestry systems in
combating land degradation and enhancing soil fertility. Incorporation of A. senegal,
in the traditional dryland agroforestry systems in Sudan, has contributed to approx-
imately 217 and 1,500 kg ha–1 of K and soil organic carbon (SOC), respectively, to
the top 25 cm of soil during the first four years of intercropping with food crops such
as sorghum and sesame [68]. Likewise, agroforestry systems in Tigray regions,
Northern Ethiopia, displayed the highest SOC associated with microaggregates
(2 × 6 g kg-1) compared to rainfed cultivation, open pasture, silvopasture, and
irrigation agriculture systems, an indication of its potential to stabilize SOC more
than the other land uses [69].

5.3 Indigenous Grass Reseeding

Reseeding using indigenous perennial grasses has been used as a strategy to combat
land degradation in arid and semi-arid drylands in Kenya (Fig. 4). Perennial grasses,
e.g., Chloris roxburghiana Schult., Cenchrus ciliaris L., Enteropogon
macrostachyus (Hochst. Ex A. Rich.) Monro ex Benth., Eragrostis superba Peyr.,
Chloris gayana Knuth., Panicum coloratum L., Sorghum sudanense (P.) Stapf., and
Cynodon dactylon (L.) Pers., are among the commonly used native species for
reseeding Kenyan drylands. Morphometric characteristics of selected perennial
grasses used for reseeding Kenyan dryland environments are shown in Table 2.
Selection of these species has been informed by their ability to tolerate drought,
produce biomass of good quantity and forage value, produce an adequate amount of
viable seed which can be easily harvested and easily established, tolerate grazing
pressure, and establish fast during spells of favorable climatic conditions. For
example, in order to take advantage of the low and sporadic rainfall in the arid and
semi-arid zones in Kenya, C. ciliaris, and C. roxburghiana have a high concentra-
tion of root biomass (about 75%) in the upper 0–30 cm soil depth [70, 71].
Agricultural Land Degradation in Kenya 291

Fig. 4 Chloris gayana cv. Boma Rhodes established in a semi-arid landscape in Kenya (Photo
credit: Kevin Z. Mganga)

Table 2 Morphometric characteristics of selected perennial grasses used for reseeding Kenyan
dryland environments
Perennial grasses native to Kenyan drylands
Cenchrus Enteropogon Eragrostis Chloris Chloris
Plant trait ciliaris macrostachyus superba roxburghiana gayana
Distribution 0–2,000 m 300–1,600 m 0–2,000 m 30–1,500 m 0–2,400 m
a.s.l. a.s.l a.s.l. a.s.l a.s.l.
Leaf blades 15–30 cm 10–60 cm long Up to 40 cm 15–30 cm 12–50 cm
long 2–10 mm wide long long long
3–8 mm 3–12 mm 3–6 mm wide 10–20 mm
wide wide wide
Stem 1–2 mm 1–2 mm wide 2–3 mm 2–3 mm wide 4–5 mm
thickness wide wide wide
Plant height 20–150 cm 30–100 cm 20–120 cm 40–150 cm 50–220 cm
Rooting Up to Up to Up to 100 cm Up to
depth 240 cm 220 cm 470 cm
Inflorescence Panicle 2– Spike 8–20 cm Panicle 10– Spike 6– Racemes 4–
14 cm long long 30 cm long 18 cm long 15 cm long
Optimal 300– 550–800 mm 500–900 mm 500–625 mm 600–750 mm
annual 750 mm
rainfall
% Crude Up to 10% 9–12% Up to 15% Up to 16% Up to 10%
protein
Source: Bogdan and Pratt [86]
292 K. Z. Mganga

Fig. 5 Trench to harvest rainwater and enhance the establishment of grasses native to Kenyan
drylands (Photo credit: Kevin Z. Mganga)

Pastoralists and agropastoralists inhabiting the Kenyan drylands are incorporating


different low-cost in situ rainwater harvesting techniques, e.g., half-moons/semi-
circular bunds, micro-catchments furrows, and trenches to enhance the establishment
of grasses [7, 72, 73] (Fig. 5). This is because when the seed stock is healthy, only
two environmental factors will stop it from germinating and establishing, namely
soil type and moisture. Half-moon-shaped micro-catchments in checkerboard pattern
are used to capture and retain rainwater. Semi-circular bunds have been used for the
rehabilitation of degraded semi-arid lands in Kenya and results demonstrated that
rangeland productivity of the degraded pasturelands could be achieved within three
growing seasons [63]. Ox-driven plows are also used to create shallow furrows prior
to reseeding. Also, shallow trenches are used for trapping soil and water and
promoting plant establishment. These minimal soil disturbance techniques have a
number of ultimate effects, e.g., break surface crust, support better root growth,
enhance germination of seeds and establishment of seedlings, reduce runoff,
enhance infiltration, and increase the capacity of soils to trap rainwater [10]. These
in situ rainwater harvesting techniques ensure that the soils retain enough water for a
prolonged period of time, thus improving the chances of the grass seeds to germinate
and establish.
Combining rainwater harvesting and native grass reseeding has been effective in
improving soil health and hydrological properties (increase water infiltration, reduce
runoff and sediment production), increasing vegetation and soil cover (plant densi-
ties, tiller densities, basal cover), increasing rangeland productivity, and replenishing
depleted seed bank (seed production) in Kenya drylands. Simulated rainfall in a
reseeded semi-arid dryland in Kenya showed that C. ciliaris, E. superba, and
E. macrostachyus significantly enhanced infiltration capacity and reduced runoff
Agricultural Land Degradation in Kenya 293

and sediment production across the phenological stages of development (seedling,


elongation, and maturity) [72]. Using low-cost grass restoration using erosion
barriers in a degraded dryland, Kimiti et al. [73] demonstrated that seeding with
C. ciliaris enhanced herbaceous cover, with vegetation patches ranging from 2.25 to
24.5 m2. Larger patches composed of higher plant densities are more effective at
trapping soil sediments and enhancing water infiltration than smaller patches, thus
promoting soil conservation and starting positive feedback toward increased vege-
tation cover [74]. Other studies in dryland environments in Kenya have also dem-
onstrated an increase in vegetation cover after reseeding with native perennial
grasses [7, 75].

5.4 Rangeland Enclosures

Enclosing rangelands for rehabilitating degraded arid and semi-arid grazing lands
involves closing off the denuded and depleted sections from livestock and wildlife
disturbance for at least 3 successive years, to allow the natural vegetation to
regenerate. This is an integrated landscape approach that offers numerous environ-
mental benefits, notably soil stabilization, enhanced hydrological cycles, soil nutri-
ents recharge and exchange, and C storage on a landscape level [76]. Previous
studies in the arid and semi-arid rangelands in Kenya have shown that enclosure
owners benefit from ecosystem services such as improved water infiltration and
retention, soil fertility, and erosion control [77, 78]. This rangeland rehabilitation
strategy and approach mimics the traditional nomadic pastoralism where designated
dry season pastures were set aside and not grazed. However, the shrinking rangeland
resource base and increased land use pressure have made the traditional annual
deferred grazing system among pastoralists challenging and increasingly difficult to
implement if not impossible. Subsequently, pastoralists in arid and semi-arid zones
in Kenya are using fences (e.g., live cacti, or cut Acacia thorn-bushes fences) to
enclose sections of their rangelands to rehabilitate and restore degraded patches.
Rangeland enclosures are increasingly becoming popular among pastoral commu-
nities for the rehabilitation of degraded arid and semi-arid lands in Africa [79]. This
is because enhanced vegetation cover in established enclosures is increasing soil
cover, thus minimizing loss of soil moisture through evapotranspiration, and enhanc-
ing soil water infiltration, thus reducing runoff and soil erosion. Moreover, deposi-
tion of plant litter and C storage have also enhanced soil fertility, hence increasing
rangeland productivity.
Approximately 38 communal enclosures covering a land area of about 1,496 ha
and well over 700 private enclosures ranging in size from 05 to 20 ha in the whole
Lake Baringo basin have been established [46]. Studies have shown that communal
and private enclosures enhance soil chemical and biological properties [77, 80,
81]. Both communal and private enclosures significantly increased organic C and
total N contents compared to open semi-arid rangelands in Lake Baringo basin. A
previous study demonstrated that organic soil C contents in an open semi-arid
294 K. Z. Mganga

rangeland increased from 3.91 to 5.60 and 3.69 to 9.20 mg C g-1 soil in the private
and communal enclosures, respectively. Similarly, total soil N contents in the same
rangeland increased from 0.41 mg N g-1 soil to 0.61 and 0.75 mg N g-1 soil in the
private and communal enclosures, respectively [77]. Soil microbial biomass C
content of 99.8 and 137.8 μg C g-1 soil in private and communal enclosures,
respectively, was significantly higher compared to 58 μg C g-1 soil recorded in
the open rangeland. Likewise, communal (61.7 μg N g-1 soil) and private
(39.5 μg N g-1 soil) enclosures had higher microbial biomass N content than open
rangeland of up to 30.0 μg N g-1 soil [77]. However, a 23-year chronosequence
analysis of communal and private enclosures in a degraded semi-arid rangeland in
Kenya demonstrated that soil health generally improved at a slow rate with time and
that the level of biochemical indicators of soil quality, e.g., soil microbial biomass,
obtained under both enclosure systems was still low [81]. Subsequently, this obser-
vation highlights the significance of continuous monitoring of livestock grazing and
forage harvesting in rangeland enclosures to ensure sustainable utilization of grazing
resources and combat degradation in arid and semi-arid zones in Kenya.
Similar studies in the Ethiopian drylands have also demonstrated that ecological
sustainability is the main reason for the establishment of rangeland enclosures
[82, 83]. In the Ethiopian drylands, enclosures are a form of diversification to an
alternative source of livelihood and a coping strategy among pastoralists against the
highly variable rangeland forage resource base. Borana pastoralists in southern
Ethiopia use enclosures as dry season grazing reserves for lactating cows, calves,
and sick and/or weak livestock. Similar to the Kenyan drylands, a high prevalence of
community enclosures and the trend toward setting up of new ones is increasing
among the Borana in southern Ethiopia [84].

6 Conclusions

Soil and land degradation characterized by soil erosion, depletion of soil nutrients,
and decline in fertility depicted by deficiencies in micro- and macronutrients limits
primary production and biological processes such as N-fixation. Despite these
inherent challenges in Kenyan arid and semi-arid drylands, sustainable land man-
agement practices have demonstrated great potential to combat soil degradation.
Conservation agriculture, dryland agroforestry, native grass reseeding, and range-
land enclosures are strategies that can halt and reverse the land degradation process
and mitigate the impacts of climate change and variability in these dryland environ-
ments. These approaches contribute to sustainable crop and livestock production
through increasing and maintaining vegetation cover, rehabilitating soil
hydrophysical properties, increasing rainwater use efficiency, and enhancing soil
fertility.
Despite their significant contribution in combating soil and land degradation,
these strategies might also pose a threat to agroecosystem goods and services. For
instance, agroforestry systems have resulted in the reduction in crop yields in
Agricultural Land Degradation in Kenya 295

Kenyan drylands. This has been attributed to the competitive advantage trees and
shrubs have over food crops for light, water, and nutrients. Such emerging chal-
lenges offer researchers and other stakeholders with an excellent opportunity to
investigate how existing and new sustainable land management practices can best be
adapted to specific local, national, and regional contexts. Ideally, the choice and
adoption of these sustainable land management practices at farm and/or agricultural
landscape level should largely be informed by the anticipated short- and long-term
benefits vis-à-vis the potential trade-offs. Thus, there is a need to appreciate and
comprehend the complexities of using sustainable land management practices to
effectively combat land degradation. To achieve this, further research should aim to
identify and develop suitable solutions that will not only prevent and reverse land
degradation but also improve crop yields and increase resilience of land-based
production systems.

References

1. Kabubo-Mariara J (2008) Climate change adaptation and livestock activity choices in Kenya: an
economic analysis. Nat Res Forum 32:131–141. https://doi.org/10.1111/j.1477-8947.2008.
00178.x
2. Amwata DA, Nyariki DM, Musimba NKR (2016) Factors influencing pastoral and agropastoral
household vulnerability to food insecurity in the drylands of Kenya: a case study of Kajiado and
Makueni counties. J Int Dev 28:771–787. https://doi.org/10.1002/jid.3123
3. Mumo L, Yu J, Ayugi B (2019) Evaluation of spatiotemporal variability of rainfall over Kenya
from 1979-2017. J Atmos Sol Terr Phys 194:105097. https://doi.org/10.1016/j.jastp.2019.
105097
4. Kimiti JM, Odee DW (2010) Integrated soil fertility management enhances population and
effectiveness of indigenous cowpea rhizobia in semi-arid eastern Kenya. Appl Soil Ecol 45:
304–309. https://doi.org/10.1016/j.apsoil.2010.05.008
5. Caretta MA, Westerberg L-O, Mburu DW, Fischer M, Börjeson L (2018) Soil management and
soil properties in a Kenyan smallholder irrigation system on naturally low-fertile soils. Appl
Geogr 90:248–256. https://doi.org/10.1016/j.apgeog.2017.12.008
6. Oba G, Nordal I, Stenseth NC, Stave J, Bjora CS, Muthondeki JK, Bii WKA (2001) Growth
performance of exotic and indigenous tree species in saline soils in Turkana, Kenya. J Arid
Environ 47:499–511. https://doi.org/10.1006/jare.2000.0734
7. Mganga KZ, Musimba NKR, Nyariki DM, Nyangito MM, Mwang’ombe AW (2015) The
choice of grass species to combat desertification in semi-arid rangelands is greatly influenced by
their forage value for livestock. Grass Forage Sci 70:161–167. https://doi.org/10.1111/gfs.
12089
8. Opiyo F, Wasonga O, Nyangito M, Schilling J, Munang R (2015) Drought adaptation and
coping strategies among the Turkana pastoralists of northern Kenya. Int J Disaster Risk Sci 6:
295–309. https://doi.org/10.1007/s13753-015-0063-4
9. Finkel M, Darkoh MBK (1991) Sustaining the arid and semi-arid (ASAL) environment in
Kenya through improved pastoralism and agriculture. J East Africa Res Dev 21:1–20. http://
www.jstor.org/stable/24326241
10. Visser N, Morris C, Hardy MB, Botha JC (2007) Restoring bare patches in the Nama-Karoo of
South Africa. Afr J Range Forage Sci 24:87–96. https://doi.org/10.2989/AJRFS.2007.24.2.
5.159
296 K. Z. Mganga

11. Prăvălie R, Bandoca G, Patriche C, Sternberg T (2019) Recent changes in global drylands:
evidences from two major aridity databases. Catena 178:209–231. https://doi.org/10.1016/j.
catena.2019.03.016
12. Huang J, Ji M, Xie Y, Wang S, He Y, Ran J (2016) Global semi-arid climate change over last
60 years. Climate Dynam 46:1131–1150. https://doi.org/10.1007/s00382-015-2636-8
13. Mganga KZ, Nyariki DM, Musimba NKR, Amwata DA (2018) Determinants and rates of land
degradation: application of stationary time-series model to data from a semi-arid environment in
Kenya. J Arid Land 10:1–11. https://doi.org/10.1007/s40333-017-0036-0
14. Lal R (2009) Sequestering carbon in soils of arid ecosystems. Land Degrad Dev 20:441–454.
https://doi.org/10.1002/ldr.934
15. Akinyemi FO, Ghazaryan G, Dubovyk O (2021) Assessing UN indicators of land degradation
neutrality and proportion of degraded land for Botswana using remote sensing based national
level metrics. Land Degrad Dev 32:158–172. https://doi.org/10.1002/ldr.3695
16. Waswa BS, Vlek PLG, Tamene LD, Okoth P, Mbakaya D, Zingore S (2013) Evaluating
indicators of land degradation in smallholder farming systems of western Kenya. Geoderma
195–196:192–200. https://doi.org/10.1016/j.geoderma.2012.11.007
17. Barbier EB (2000) The economic linkages between rural poverty and land degradation: some
evidence from Africa. Agric Ecosyst Environ 82:355–370. https://doi.org/10.1016/S0167-8809
(00)00237-1
18. Requier-Desjardins M, Adhikari B, Sperlich S (2011) Some notes on the economic assessment
of land degradation. Land Degrad Dev 22:285–298. https://doi.org/10.1002/ldr.1056
19. Kassahun A, Snyman HA, Smit GN (2008) Livestock grazing behaviour along a degradation
gradient in the Somali region of eastern Ethiopia. Afr J Range Forage Sci 25:1–9. https://doi.
org/10.2989/AJRFS.2008.25.1.1.379
20. Mulinge W et al (2016) Economics of land degradation and improvement in Kenya. In:
Nkonya E, Mirzabaev A, von Braun J (eds) Economics of land degradation and improvement –
a global assessment for sustainable development. Springer, Cham, pp 471–498. https://doi.org/
10.1007/978-3-319-19168-3_16
21. Dregne HE (2002) Land degradation in the drylands. Arid Land Res Manag 16:99–132. https://
doi.org/10.1080/153249802317304422
22. Butt B (2010) Pastoral resource access and utilization: quantifying the spatial and temporal
relationships between livestock mobility, density and biomass availability in southern Kenya.
Land Degrad Dev 21:520–539. https://doi.org/10.1002/ldr.989
23. Boles OJC, Shoemaker A, Mustaphi CJC, Petek N, Eklom A, Lane PJ (2019) Historical
ecologies of pastoralist overgrazing in Kenya: long-term perspectives on cause and effect.
Hum Ecol 47:419–434. https://doi.org/10.1007/s10745-019-0072-9
24. Eschen R, Bekele K, Mbaabu PR, Kilawe CJ, Eckert S (2021) Prosopis juliflora management
and grassland restoration in Baringo County, Kenya: opportunities for soil carbon sequestration
and local livelihoods. J Appl Ecol 00:1–12. https://doi.org/10.1111/1365-2664.13854
25. Kumar R, Chandrashekar N (2016) Study on fuelwood and carbonization characteristics of
Prosopis juliflora. J Indian Acad Wood Sci 13:101–107. https://doi.org/10.1007/s13196-016-
0171-9
26. Fagg CW, Stewart JL (1994) The value of Acacia and Prosopis in arid and semi-arid environ-
ments. J Arid Environ 27:3–25. https://doi.org/10.1006/jare.1994.1041
27. Eberhard AA (1990) Fuel calorific values in South Africa. South Afr For J 152:17–22. https://
doi.org/10.1080/00382167.1990.9629014
28. Njenga M, Karanja N, Munster C, Iiyama M, Neufeldt H, Kithinji J, Jamnadass R (2013)
Charcoal production and strategies to enhance its sustainability in Kenya. Dev Pract 23:359–
371. https://doi.org/10.1080/09614524.2013.780529
29. Muchena FN, Onduru DD, Gachini GN, de Jager A (2005) Turning the tides of soil degradation
in Africa: capturing the reality and exploring opportunities. Land Use Policy 22:23–31. https://
doi.org/10.1016/j.landusepol.2003.07.001
Agricultural Land Degradation in Kenya 297

30. Onduru DD, de Jager A, Gachini GN (2005) The hidden costs of soil mining to agricultural
sustainability in developing countries: a case study of Machakos District, Eastern Kenya. Int J
Agric Sustain 3:167–176. https://doi.org/10.1080/14735903.2005.9684754
31. Gachimbi LN, van Keulen H, Thuranira EG, Karuku AM, de Jager A, Nguluu S, Ikombo BM,
Kinama JM, Itabari JK, Nandwa SM (2005) Nutrient balances at farm level in Machakos
(Kenya), using a participatory nutrient monitoring (NUTMON) approach. Land Use Policy
22:13–22. https://doi.org/10.1016/j.landusepol.2003.07.002
32. Onyando JO, Kisoyan P, Chemelil MC (2005) Estimation of potential soil erosion for River
Perkerra catchment in Kenya. Water Resour Manag 19:133–143. https://doi.org/10.1007/
s11269-005-2706-5
33. Defersha MB, Melesse AM, McClain ME (2012) Watershed scale application of WEPP and
EROSION 3D models for assessment of potential sediment source areas and runoff flux in the
Mara River Basin, Kenya. Catena 95:63–72. https://doi.org/10.1016/j.catena.2012.03.004
34. Nkedianye D, de Leeuw J, Ogutu JO, Said MY, Saidimu TL, Kifugo SC, Kaelo DS, Reid RS
(2011) Mobility and livestock mortality in communally used pastoral areas: the impact of the
2005–2006 droughts on livestock mortality in Maasailand. Pastoralism: Res Policy Pract 1:1–
17. https://doi.org/10.1186/2041-7136-1-17
35. Kabubo-Mariara J (2009) Global warming and livestock husbandry in Kenya: impacts and
adaptations. Ecol Econ 68:1915–1924. https://doi.org/10.1016/j.ecolecon.2009.03.002
36. Xie H, Perez N, Anderson W, Ringler C, You L (2018) Can Sub-Saharan Africa feed itself? The
role of irrigation development in the region’s drylands for food security. Water Int 43:796–814.
https://doi.org/10.1080/02508060.2018.1516080
37. Farah KO, Nyariki DM, Noor AA, Ngugi RK, Musimba NK (2003) The socio-economic and
ecological impacts of small-scale irrigation schemes on pastoralists and drylands in Northern
Kenya. J Soc Sci 7:267–274. https://doi.org/10.1080/09718923.2003.11892389
38. Macharia DW, Muriuki SK (1986) Changes in saturated hydraulic conductivity in selected soils
of Kenya irrigated with waters of variable salt content. East Afr Agric For J 52:47–55. https://
doi.org/10.1080/00128325.1986.11663493
39. Tiffen M, Mortimore M, Gichuki F (1994) More people, less erosion: environmental recovery
in Kenya. Overseas Development Institute, Wiley, London
40. Hilson G (2016) Farming, small-scale mining and rural livelihoods in Sub-Saharan Africa: a
critical overview. Ext Ind Soc 3:547–563. https://doi.org/10.1016/j.exis.2016.02.003
41. Apollo F, Ndinya A, Ogada M, Rop B (2019) Feasibility and acceptability of environmental
management strategies among artisan miners in Taita Taveta County, Kenya. J Sustain Min 16:
189–195. https://doi.org/10.1016/j.jsm.2017.12.003
42. Mwakumanya MA, Maghenda M, Juma H (2016) Socio-economic and environmental impact
of mining on women in Kasigau mining zone in Taita Taveta County. J Sustain Min 15:197–
204. https://doi.org/10.1016/j.jsm.2017.04.001
43. Muriuki GW, Njoka TJ, Reid RS, Nyariki DM (2005) Tsetse control and land-use change in
Lambwe valley, south-western Kenya. Agric Ecosyst Environ 106:99–107. https://doi.org/10.
1016/j.agee.2004.04.005
44. Burian A, Karaya R, Wernersson JEV, Egberth M, Lokorwa B, Nyberg G (2019) A community-
based evaluation of population growth and agro-pastoralist resilience in Sub-Saharan drylands.
Environ Sci Policy 92:323–330. https://doi.org/10.1016/j.envsci.2018.10.021
45. Martin DM (2017) Ecological restoration should be redefined for the twenty-first century.
Restor Ecol 25:668–673. https://doi.org/10.1111/rec.12554
46. Mureithi SM, Verdoodt A, Njoka JT, Gachene CKK, van Ranst E (2016) Benefits derived from
rehabilitating a degraded semi-arid rangeland in communal enclosures, Kenya. Land Degrad
Dev 27:1853–1862. https://doi.org/10.1002/ldr.2341
47. Verdoodt A, Mureithi SM, van Ranst E (2010) Impacts of management and enclosure age on
recovery of the herbaceous rangeland vegetation in semi-arid Kenya. J Arid Environ 74:1066–
1073. https://doi.org/10.1016/j.jaridenv.2010.03.007
298 K. Z. Mganga

48. Osborne M (2014) Controlling development: ‘Martial Race’ and Empire in Kenya, 1945-59. J
Imp Commonw Hist 42(3):464–485. https://doi.org/10.1080/03086534.2013.868230
49. Palm C, Blanco-Canqui H, DeClerck F, Gatere L, Grace P (2014) Conservation agriculture and
ecosystem services: an overview. Agric Ecosyst Environ 187:87–105. https://doi.org/10.1016/j.
agee.2013.10.010
50. Six J, Elliott ET, Paustian K (2000) Soil macroaggregate turnover and microaggregate forma-
tion: a mechanism for C sequestration under no-tillage agriculture. Soil Biol Biochem 32:2099–
2103. https://doi.org/10.1016/S0038-0717(00)00179-6
51. Kihara J, Bationo A, Mugendi DN, Martius C, Vlek PLG (2011) Conservation tillage, local
organic resources and nitrogen fertilizer combinations affect maize productivity, soil structure
and nutrient balances in semi-arid Kenya. Nutr Cycl Agroecosyst 90:213–225. https://doi.org/
10.1007/s10705-011-9423-7
52. Mutuku EA, Roobroeck VB, Boeckx P, Cornelis WM (2020) Maize production under com-
bined Conservation Agriculture and Integrated Soil Fertility Management in the sub-humid and
semi-arid regions of Kenya. Field Crop Res 254:107833. https://doi.org/10.1016/j.fcr.2020.
107833
53. Thierfelder C, Wall PC (2009) Effects of conservation agriculture techniques on infiltration and
soil water content in Zambia and Zimbabwe. Soil Tillage Res 105:217–227. https://doi.org/10.
1016/j.still.2009.07.007
54. Snapp SS, Blackie MJ, Gilbert RA, Bezner-Kerr R, Kanyama-Phiri GY (2010) Biodiversity can
support a greener revolution in Africa. Proc Natl Acad Sci 107:20840–20845. https://doi.org/
10.1073/pnas.1007199107
55. Rao MR, Mathuva MN (2000) Legumes for improving maize yields and income in semi-arid
Kenya. Agric Ecosyst Environ 78:123–137. https://doi.org/10.1016/S0167-8809(99)00125-5
56. Kiwia A, Kimani D, Harawa R, Jama B, Sileshi GW (2019) Sustainable Intensification with
Cereal-Legume Intercropping in Eastern and Southern Africa. Sustainability 11:2891. https://
doi.org/10.3390/su11102891
57. Myaka FM, Sakala WD, Adu-Gyamfi JJ, Kamalongo D, Ngwira A, Odgaard R, Nielsen NE,
Høgh-Jensen H (2006) Yields and accumulations of N and P in farmer-managed intercrops of
maize–pigeonpea in semi-arid Africa. Plant and Soil 285:207–220. https://doi.org/10.1007/
s11104-006-9006-6
58. Kwena K, Karuku GN, Ayuke FO, Esilaba AO (2019) Nitrogen deficiency in semi-arid Kenya:
can Pigeonpea fix it? East Afr Agric For J 83:322–340. https://doi.org/10.1080/00128325.2019.
1658696
59. Wood M, McNeill AM, Pilbeam CJ, Swift RS, Harris HC, Mugane PG (1996) Sustainability of
nitrogen use in two dryland farming systems. In: Van Cleemput O, Hofman G, Vermoesen A
(eds) Progress in nitrogen cycling studies. Developments in plant and soil sciences. Springer,
Dordrecht, pp 303–306. https://doi.org/10.1007/978-94-011-5450-5_51
60. Smith RG, Gross KL, Robertson GP (2008) Effects of crop diversity on agroecosystem
function: crop yield response. Ecosystems 11:355–366. https://doi.org/10.1007/s10021-008-
9124-5
61. Chivenge P, Vanlauwe B, Gentile R, Wangechi H, Mugendi D, van Kessel C, Six J (2009)
Organic and mineral input management to enhance crop productivity in central Kenya. Agron J
101:1266–1275. https://doi.org/10.2134/agronj2008.0188x
62. Chivenge PP, Murwira HK, Giller KE, Mapfumo P, Six J (2007) Long-term impact of reduced
tillage and residue management on soil carbon stabilization: implications for conservation
agriculture on contrasting soils. Soil Tillage Res 94:328–337. https://doi.org/10.1016/j.still.
2006.08.006
63. Itabari JK, Kwena K, Esilaba AO, Kathuku AN, Muhammad L, Mangale N, Kathuli P (2011)
Land and water management research and development in arid and semi-arid lands of
Kenya. In: Bationo A, Waswa B, Okeyo J, Maina F, Kihara J (eds) Innovations as key to the
Green Revolution in Africa. Springer, Dordrecht, pp 427–438. https://doi.org/10.1007/978-90-
481-2543-2_44
Agricultural Land Degradation in Kenya 299

64. Cooper PJM, Leakey RRD, Rao MR, Reynolds L (1996) Agroforestry and the mitigation of
land degradation in the humid and sub-humid tropics of Africa. Exp Agric 32:235–290. https://
doi.org/10.1017/S0014479700026223
65. Ndegwa G, Iiyama M, Anhuf D, Nehren U, Schlüter S (2017) Tree establishment and manage-
ment on farms in the drylands: evaluation of different systems adopted by small-scale farmers in
Mutomo District, Kenya. Agrofor Syst 91:1043–1055. https://doi.org/10.1007/s10457-016-
9979-y
66. Kinama JM, Stigter CJ, Ong CK, Ng'ang'a JK, Gichuki FN (2007) Contour hedgerows and
grass strips in erosion and runoff control on sloping land in semi-arid Kenya. Arid Land Res
Manag 21:1–19. https://doi.org/10.1080/15324980601074545
67. Lehmann J, Weigl D, Droppelmann K, Huwe B, Zech W (1999) Nutrient cycling in an
agroforestry system with runoff irrigation in Northern Kenya. Agr Syst 43:49–70. https://doi.
org/10.1023/A:1026447119829
68. Raddad EY, Luukkanen O, Salih AA, Kaarakka V, Elfadl MA (2006) Productivity and nutrient
cycling in young Acacia senegal farming systems on Vertisol in the Blue Nile region, Sudan.
Agr Syst 68:193–207. https://doi.org/10.1007/s10457-006-9009-6
69. Gelaw AM, Singh BR, Lal R (2015) Organic carbon and nitrogen associated with soil
aggregates and particle sizes under different land uses in Tigray, Northern Ethiopia. Land
Degrad Dev 26:690–700. https://doi.org/10.1002/ldr.2261
70. Marshall VM, Lewis MM, Ostendorf B (2012) Buffel grass (Cenchrus ciliaris) as an invader
and threat to biodiversity in arid environments: a review. J Arid Environ 78:1–12. https://doi.
org/10.1016/j.jaridenv.2011.11.005
71. Macharia PN, Kinyamario JI, Ekaya WN, Gachene CKK, Mureithi JG, Thuranira EG (2010)
Evaluation of forage legumes for introduction into natural pastures of semi-arid rangelands of
Kenya. Grass Forage Sci 65:456–462. https://doi.org/10.1111/j.1365-2494.2010.00764.x
72. Mganga KZ, Nyariki DM, Musimba NKR, Mwang’ombe AW (2019) Indigenous grasses for
rehabilitating degraded African drylands. In: Bamutaze Y, Kyamanywa S, Singh B,
Nabanoga G, Lal R (eds) Agriculture and ecosystem resilience in Sub Saharan Africa. Climate
Change Management. Springer, Cham, pp 53–68. https://doi.org/10.1007/978-3-030-12974-3_
3
73. Kimiti DW, Riginos C, Belnap J (2017) Low-cost grass restoration using erosion barriers in a
degraded African rangeland. Restor Ecol 25:376–384. https://doi.org/10.1111/rec.12426
74. Ludwig JA, Wiens JA, Tongway DJ (2000) A scaling rule for landscape patches and how it
applies to conserving soil resources in savannas. Ecosystems 3:84–97. https://doi.org/10.1007/
s100210000012
75. Kinyua D, McGeoch LE, Georgiadis N, Truman PY (2010) Short-term and long-term effects of
soil ripping, seeding and fertilisation on the restoration of a tropical rangeland. Restor Ecol 18:
226–233. https://doi.org/10.1111/j.1526-100X.2009.00594.x
76. Scherr SJ, Shames S, Friedman R (2012) From climate-smart agriculture to climate-smart
landscapes. Agric Food Secur 1:12. https://doi.org/10.1186/2048-7010-1-12
77. Mureithi SM, Verdoodt A, Gachene CKK, Njoka JT, Wasonga VO, de Neve S, Meyerhoff E,
van Ranst E (2014) Impacts of enclosure management on soil properties and microbial biomass
in a restored semi-arid rangeland, Kenya. J Arid Land 6:561–570. https://doi.org/10.1007/
s40333-014-0065-x
78. Nyberg G, Knutsson P, Ostwald M et al (2015) Enclosures in West Pokot, Kenya: transforming
land, livestock and livelihoods in drylands. Pastoralism 5:25. https://doi.org/10.1186/s13570-
015-0044-7
79. Wairore JN, Mureithi SM, Wasonga OV, Nyberg G (2015) Enclosing the commons: reasons for
the adoption and adaptation of enclosures in the arid and semi-arid rangelands of Chepareria,
Kenya. SpringerPlus 4:595. https://doi.org/10.1186/s40064-015-1390-z
80. Feyisa K, Beyene S, Angassa A, Said MY, de Leeuw J, Abebe A, Megersa B (2017) Effects of
enclosure management on carbon sequestration, soil properties and vegetation attributes in East
African rangelands. Catena 159:9–19. https://doi.org/10.1016/j.catena.2017.08.002
300 K. Z. Mganga

81. Verdoodt A, Mureithi SM, Ye L, van Ranst E (2009) Chronosequence analysis of two enclosure
management strategies in degraded rangeland of semi-arid Kenya. Agric Ecosyst Environ 129:
332–339. https://doi.org/10.1016/j.agee.2008.10.006
82. Angassa A, Oba G (2010) Effects of grazing pressure, age of enclosures and seasonality on bush
cover dynamics and vegetation composition in southern Ethiopia. J Arid Environ 74:111–120.
https://doi.org/10.1016/j.jaridenv.2009.07.015
83. Fenetahun Y, Yuan Y, Xinwen X, Yongdong W (2021) Effects of grazing enclosures on species
diversity, phenology, biomass, and carrying capacity in Borana Rangeland, Southern Ethiopia.
Front Ecol Evol 8:623627. https://doi.org/10.3389/fevo.2020.623627
84. Angassa A, Oba G (2008) Herder perceptions on impacts of range enclosures, crop farming, fire
ban and bush encroachment on the Rangelands of Borana, Southern Ethiopia. Hum Ecol 36:
201–215. https://doi.org/10.1007/s10745-007-9156-z
85. Sombroek WG, Braun HMH, van der Pouw BJA (1982) The exploratory soil map and agro-
climate zone map of Kenya (1980) scale 1:1,000,000. Exploratory Soil Survey Report E1,
Kenya Soil Survey, Nairobi
86. Bogdan AV, Pratt DJ (1967) Reseeding denuded pastoral land in Kenya. Republic of Kenya,
Ministry of Agriculture and Animal Husbandry, Government Printers, pp 1–46
Agricultural Land Degradation in Mexico

Nadia S. Santini, Angela P. Cuervo-Robayo, and María Fernanda Adame

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
2 Agriculture in Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
2.1 Traditional Farming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
2.2 Intensive Industrial Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
3 The Soils of Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
4 The Health of Mexican Soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
4.1 Physical Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
4.2 Chemical Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
4.3 Water and Wind Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
5 Examples of Soil Degradation Across Mexican Ecoregions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
5.1 Mexican Temperate Highlands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
5.2 Tropical and Subtropical Dry Forests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
5.3 Degradation of North American Deserts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
6 Sustainable Agriculture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319

Abstract Mexico is one of the most diverse countries on Earth and is a centre of
origin for many major global crops, including bean, cotton, chilli, pumpkin, and
avocado. During the past decades, a growing demand for food has promoted clearing

N. S. Santini (✉)
Cátedra Consejo Nacional de Ciencia y Tecnología, Ciudad de México, México
Instituto de Ecología, Universidad Nacional Autónoma de México, Ciudad de México, México
e-mail: [email protected]
A. P. Cuervo-Robayo
Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, Ciudad de México,
Mexico
M. F. Adame
Australian Rivers Institute, Griffith University, Nathan, QLD, Australia
e-mail: f.adame@griffith.edu.au

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 301
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 301–324, DOI 10.1007/698_2022_915,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 6 December 2022
302 N. S. Santini et al.

and conversion of Mexican forests to arable land, the irrigation of arid regions using
limited water resources and the transition of traditional farming practices to more
intensive agricultural practices. This rapid conversion of natural ecosystems and
poor agricultural practices have caused a marked decline in the soil health, including
physical and chemical degradation as well as water and wind erosion. This deteri-
oration in soil health results in a loss of microbial biodiversity, organic and mineral
nutrients and water. Here, we describe how different agricultural practices have
impacted the range of soil types across the diverse Mexican ecoregions. Despite the
negative impact, the Mexican government has proposed to expand and intensify
agriculture to increase crop output. Unless new policy regulations, such as the
supervision of chemical and fertilizer use, are urgently applied, this expansion and
intensification in agriculture will further reduce soil health, resulting in further losses
to biodiversity and environmental services upon which Mexican agriculture ulti-
mately depends.

Keywords Avocado plantations, Conservation agriculture, Cuatro Ciénegas,


Mexican ecoregions, Soil degradation, Yucatán Peninsula

1 Introduction

Soils are one Earth’s essential resource and provide a substrate for plant growth,
storage of water and nutrients and a habitat for a vast diversity of organisms. Soils
provide critical services to humans, including supporting the agricultural production
of crops and food. However, soil is a finite resource, and its structure, nutrients and
organisms and biodiversity can be lost [1]. Accordingly, the sustainable manage-
ment of soil health is necessary to assure these essential environmental services, that
are needed for human well-being and prosperity.
The health of soil is the result of physical, chemical and biological properties.
Physical properties include the texture and structure of the soil. Healthy soils are
porous, allowing the exchange of gases and water with the environment and the
movement of organisms and the growth of plant roots [2, 3]. The chemical properties
of soil include characteristics such as soil pH and cation exchange capacity, impor-
tant for the retention of nutrients, with carbon, nitrogen and phosphorus especially
important to support plant growth [4, 5]. Finally, the biological properties of a
healthy soil include the diverse range of biota that reside within the soil, including
bacteria, fungi, protozoa, nematodes, earthworms and arthropods. Soil biota has
many key roles in soil health, including degrading and releasing nutrients from
organic detritus, converting pesticides and improving soil structure [3, 6].
Agricultural practices have indelibly altered the landscape of Mexico. Between
2001 and 2018, Mexico lost 42,785 ha of primary forest per year due to conversion
to agricultural land. An additional 157,528 ha per year of forests was converted to
grasslands for cattle ranching and 6,035 ha per year to human settlements [7]. As
Agricultural Land Degradation in Mexico 303

expected, this conversion of natural ecosystem has also had a major impact on soil
composition and health.
A decline in soil health can result in the loss of biodiversity, nutrients and water
and limits the environmental services provided by soil, e.g. [8, 9]. Soil degradation
can result from poor land management, soil compaction, erosion and pollution
[10]. Unsustainable agricultural practices, deforestation and losses of vegetative
cover can intensify erosion, whilst fertilizers and pesticides can pollute the soils
and groundwater [11]. Here, we review the type and health of Mexican soils, and the
impact of agriculture and pastoral land use upon them. We provide examples from
across Mexico that demonstrate how unsustainable agricultural practices can lead to
physical and chemical degradation and erosion and discuss sustainable strategies for
managing the health of Mexican soils.

2 Agriculture in Mexico

Mexico is one of the cradles of agriculture, and agriculture continues to provide a


major contribution to Mexican society and economy. In 2016, Mexico produced
146 million tons of agricultural products, earning US$18.5 billion dollars
(corresponding to 2.3% to the Gross Domestic Product) [12]. Furthermore, this
agricultural production is expected to continue to grow by 28% to 2030 and remains
a primary contributor to the Mexican economy.
Mexico is one of the world’s most biodiverse countries and encompasses seven
major distinct terrestrial ecoregions, from tropical forest to arid deserts. It is a centre
of origin and genetic diversity for many globally important crops, including bean
(Phaseolus spp.), cotton (Gossypium spp.), chilli (Capsicum spp.), pumpkin
(Cucurbita spp.), Mexican husk tomato or tomatillo (Physalis philadelphica),
vanilla (Vanilla spp.), maguey (Agave spp.) and avocado (Persea spp.) Mexico is
also one of the world’s largest producers of tomatoes (Lycopersicum esculentum),
chillies, berries (Fragaria L, Vaccinium spp., Rubus ulmifolius, Rubus idaeus),
avocado and agave (Agave tequilana) [12].
Arguably, the most important crop in Mexico is maize (Zea mays), which
occupies 28% (7.5 million ha) of the cultivated land area across Mexico. Mexico
is the centre of origin for maize, which has also assumed cultural and economic
significance throughout Mexican history [13]. As a staple crop, maize ensures the
food security of millions of families, with nearly 2.8 million farmers cultivating this
crop and contributing 18% of the agricultural sector’s production value, which
corresponded to US$5 billion dollars in 2018 [14].

2.1 Traditional Farming

Throughout Mexican history, traditional farming practices have been used to pro-
duce agricultural goods. Over 80% of Mexico’s agricultural land has been
304 N. S. Santini et al.

Fig. 1 Photographs showing examples of soils, farming practices and ecoregions throughout
Mexico. (a) Traditional agriculture in the tropical and subtropical dry forests of Chiapas, Mexico.
Tzotzil indigenous groups use crop rotation of maize, pumpkin and bean as well as other crops such
as chilli and soursop. (b) Expansion of maize fields results in a mosaic landscape of secondary
vegetation in Chiapas. (c) Land abandonment in the tropical and subtropical dry forests of Yucatán
during the rainy season, when shrubby secondary vegetation flourishes. (d) Conversion to agricul-
tural land in the Northern deserts of Coahuila, Mexico. (e) Conversion of temperate forests to
avocado orchards in Michoacán, Mexico. (f) Conversion of tropical and subtropical dry forests to
peanut and papaya farms within the Oaxacan coast. (g) Wood harvesting and deforestation in
Yucatán, Mexico. (h) Natural vegetation of the Cuatro Ciénegas Flora and Fauna Conservation
Area. (i) Intensive alfalfa farms in the Cuatro Ciénegas Basin. Photographs courtesy of Giezi
Anthony Gálvez (a, b); Anaitzi Rivero-Villar (c, g); Frédérique Reverchon (e); Nadia S. Santini
(f) and Pamela Chávez-Ortíz (d, h, i)

administered by small-scale farmers, or ‘campesinos’, that manage family farms that


primarily produce agricultural crops for local consumption (Fig. 1a). These tradi-
tional farms often rely largely on family labour and combinations of animal and
mechanical tractors, manure and organic fertilizers, and mostly employ native plant
crop varieties. Small-scale farms also typically plant and rotate a diversity of crops,
thereby promoting and conserving biodiversity, reducing tillage, and selecting seeds
that are tolerant to droughts and pests [15, 16].
Mexican farmers or campesinos have traditionally cultivated their crops in small
heterogeneous plots at different elevations or with diverse microhabitats. By
selecting and breeding crop varieties that best adapt to this diverse range of condi-
tions, Mexican farmers have increased the genetic diversity of crops. For example,
Agricultural Land Degradation in Mexico 305

campesinos have selected maize varieties to grow from sea level to highlands and
from tropical humid environments to semi-arid conditions, resulting in the diver-
gence of at least 59 different maize-landraces [17]. Traditionally, Mexican farmers
also store and share seeds among themselves, and by using and planting seeds from
heterogeneous sources, campesinos secure a yield even under adverse environmental
conditions. These traditional practices, that result in differing genetic crop varieties,
may also ensure future food security, particularly when considering uncertain future
environmental changes [18].
The impact of traditional campesino farms on soil health has been mixed.
Traditional farming typically uses crop rotation to maintain nutrients within soils
and to keep soils healthy, as well as using of local pollination and pest controls
[19]. However, traditional farming has historically used slash-and-burn clearing to
convert natural lands to agriculture. Slash-and-burn practices consist of clearing
small land areas with fire, after which the burnt area is farmed for 2–3 more years,
and then abandoned for one to two more decades (Fig. 1b, c) [20].
Slash-and-burn clearance, followed by intensive tillage can quickly drain soil
health. The deforestation and clearance of vegetation disrupts soil structure, partic-
ularly the most fertile top 20 cm, and exposes soils to wind and water erosion. The
exposure of topsoil can result in the loss of water and nutrients and the release of
greenhouse gases during burning. This practice also initiates over 40% of the bush
fires across Mexico [12]. Given the negative impacts of traditional slash-and-burn
clearing on the Mexican forests and soils, the Secretariat of Agriculture and Rural
Development (SADER by its Spanish acronym) launched a program called ‘Mi
parcela no se quema’ (in English ‘My agricultural land does not burn’) to promote
alternative practices among campesinos in 2020. These practices include retaining
the vegetative remnants after the harvest, and as such maintaining soil moisture,
nutrients and carbon within soils.

2.2 Intensive Industrial Agriculture

Intensive industrial agriculture typically uses large-scale faming, mechanized equip-


ment and large amounts of herbicides and pesticides to maximize the efficiency and
output of crop production. Over the past three decades, Mexico has prioritized the
development of intensive farming to increase crop output and re-orientate farming
towards higher-value export crops [12, 21]. This industrialization of Mexican
agriculture has been accelerated by economic liberalization, including the negotia-
tion of international trade agreements like the General Agreement on Tariffs and
Trade (GATT) in 1986 and the North American Free Trade Agreement (NAFTA) in
1994. However, this increasing industrialization of farming practices can marginal-
ize small and medium-scale farming and causes economic, environmental and social
impacts.
Industrial farming has expanded in the states of Chihuahua, Sonora and Sinaloa,
in the Yucatán Peninsula, Chiapas, Michoacán and Jalisco [7]. For example, in
306 N. S. Santini et al.

Chihuahua, the Rio Conchos watershed has been extensively modified for irrigated
cultivation of cotton, pecan nuts and sorghum. These crops were strategically
selected for domestic and international markets by the National Agricultural Pro-
gram 2017–2030 and are farmed on large-scale mechanized farms. Mexico is now
the second largest exporter of pecan nuts, mostly to China, whilst cotton production,
which has doubled in the Sonoran Desert and quadrupled in the Chihuahuan Desert,
is largely exported to the United States of America (USA) and countries from the
European Union. The increase in sorghum production is the result of a national
program to promote bioethanol making [12], and an increase in demand for fodder
for cattle raised in other regions of Mexico [22, 23].
This expansion of industrial farming in Mexico over the past decades has driven
increasing demand for agricultural land, with large areas of primary forests and arid
and semi-arid grasslands and shrublands cleared for farming in northern Mexico
(Fig. 1d). This loss of the vegetative cover is a primary cause of soil erosion in these
regions [7, 24] whilst the extensive irrigation of crops in arid and semi-arid Mexico
has also led to salinization of large dryland areas [25, 26]. Due to the use of heavy
mechanized machinery, intensive agriculture also leads to soil compaction, and soils
with high bulk density, impeding root and plant growth and resulting in lower yields.
Intensive agriculture has also caused chemical degradation of soils with extensive
and often unregulated use of chemical fertilizers and herbicides. For example,
Mexico only recently started to decrease the use of glyphosate in farming, despite
the known impact of this herbicide on soil and human health, e.g. [27]. Glyphosate
has long residence times within soils and disrupts soil nutrient dynamics and soil
microbial communities. In northern Mexico, glyphosate use has caused a decrease in
cyanobacteria from soil [24]. Given that cyanobacteria fix nitrogen and promote
plant root growth, the removal of these microbes has required farmers to increase the
use of fertilizers. Other pollutants, such as organophosphates, are also continuously
added to soils in farms and can be carried through run-off to pollute groundwater
reservoirs [28]. For example, the Yucatán Peninsula coastal water and groundwater
has been found to be polluted with glyphosate and other herbicides from
farming [27].

3 The Soils of Mexico

Mexico straddles the Pacific and North American tectonic plates and has a complex,
mountainous topography that harbours a diversity of environments. The major
mountain ranges, including the Sierra Madre Occidental and the Sierra Madre
Oriental, extend in parallel down the spine of the country, generating a diversity of
landscapes, rock and soil types across the coastal plains and interior of the country,
whilst the Yucatán Peninsula has a unique Karst limestone geography [29].
This diverse landscape has resulted in a range of different soils being developed.
The major soil types in Mexico include Leptosols (28.3% of the territory), Regosols
(13.7%), Phaeozems (11.7%), Calcisols (10.4%), Luvisols (9%) and Vertisols
Agricultural Land Degradation in Mexico 307

Fig. 2 Soil types of Mexico and degree of degradation by state. (a) The major soil types in Mexico
include Leptosols (28.3% of the territory), Regosols (13.7%), Phaeozems (11.7%), Calcisols
(10.4%), Luvisols (9%) and Vertisols (8.6%). (b) Degree of degradation across Mexican states.
(c) Degraded soil across Mexican ecoregions [30–32]

(8.6%) (Fig. 2a [30]). Much (52.4%) of the country is covered by shallow and poorly
developed Leptosols, Regosols and Calcisols (20 million ha) that are vulnerable to
erosion and degradation. Alternatively, Phaeozems, Vertisols and Luvisols are more
fertile, but are threatened by intensive agricultural use.
The Leptosols, which are common across the Sierra Madre Occidental and the
Sierra Madre Oriental mountain slopes, as well as the karstic environment of the
Yucatán Peninsula [33, 34], have a shallow profile (typically less of 25 cm depth),
contain large amount of rocks and are typically poorly developed, with low organic
308 N. S. Santini et al.

and nutrient content. Given their poor development, these soils are vulnerable to
rapid degradation, erosion and desiccation.
Regosols are a heterogeneous group of soils that are distributed throughout
Mexico, but particularly predominant in Baja California and the Sierra Madre del
Sur. Regosols are also poorly developed, with large rock and mineral content and
low clay content. These soils have a low capacity to retain water and organic matter.
Calcisols occur within the arid and semi-arid regions of Mexico and are rich in
secondary carbonates. Given the large amounts of rock fragments, dry content and
the shallow profile, these soils are vulnerable to erosion and are not typically suitable
for agriculture. However, in Mexico these soils have been used for agricultural
activities throughout intensive irrigation [33, 35].
Phaeozems soils develop from unconsolidated sediment in humid tropical and
temperate conditions, where plant growth promotes the accumulation of organic
matter within the top soil layers. Phaeozems are highly used in rainfed agriculture,
which can result in water and wind erosion [33, 35]. Luvisols form from unconsol-
idated materials and are associated with temperate or tropical climates, such as the
highlands from Oaxaca and the humid tropical forests of Chiapas. Most Luvisols in
Mexico are rich in carbon and exhibit low contents of calcium and magnesium and
are largely used for planting forage crops [35]. Vertisols occur across the fertile Gulf
Coast of Mexico and cover patches in Tamaulipas, Sonora and Sinaloa. These soils
are also common in the grasslands of Campeche and Tabasco. Vertisols are over
25 cm deep and exhibit over 30% of clay content. Vertisols are one of the most
productive agricultural soils in Mexico as they can retain water and are characterized
by a high fertility [36].
Andosols only cover 1.3% of the Mexican land, however, these soils are impor-
tant in Mexico because they originate from volcanic ashes. This material rapidly
forms secondary minerals, such as allophane or Al-humic complexes, that form
stable compounds with organic matter thereby protecting it from degradation.
Andosols are mostly distributed throughout the Trans-Mexican Volcanic Belt and
are widely exploited for agricultural activities and forestry [35, 37].

4 The Health of Mexican Soils

A healthy soil is required to provide key environmental services and support a


productive agriculture. A deterioration in soil health results in the leaching of
nutrients, poor water and gas cycling and reduced soil biodiversity, which together
reduces agricultural crop output. Despite the importance of soil health, at least 64%
of Mexican soils exhibit some degree of degradation (Fig. 2b, c). There are four
primary causes of soil degradation, including physical and chemical degradation as
well as water and wind erosion, which we describe in detail below (Fig. 3a, b).
Agricultural Land Degradation in Mexico 309

Fig. 3 Type of degradation across Mexican states and ecoregions. (a) Type of degradation across
Mexican states [30, 32]. (b) Type of degradation across Mexican ecoregions [30, 31]

4.1 Physical Degradation

The physical properties of soil include texture, which refers to the size of the soil
particles, and structure, the manner in which soil particles are aggregated. These
properties are partly a result of soil particle size (rock, sand, silt and clay), as well as
organic components and microbial activity that can aggregate together soil
particles [5].
A healthy soil that supports plant growth is usually porous, allowing an exchange
of gases with the atmosphere, and the ability of water to infiltrate and drain freely
through the soil, whilst resisting erosion. However, soil physical degradation by
compaction results in poor water filtration and drainage and reduced water retention
with greater run-off [2, 5].
Approximately 6% of the Mexican territory (10.9 million ha) is impacted by some
degree of physical degradation (Fig. 3a, b). In Mexico, physical degradation due to
soil compaction is largely due to high traffic from grazing animals and mechanized
tractors that compress the soil [31]. For example, compaction, flooding and land
subsidence are apparent in the Gulf Coastal Plain of Mexico. In this Coastal Plain,
the natural forests have been converted to grasslands for cattle ranching and lands for
310 N. S. Santini et al.

sugarcane production. As a result, Vertisols have lost their capacity to retain water
and have become highly eroded [36, 38].
Tillage refers to the digging, stirring and overturning of soil to reduce compaction
and is often used in agriculture to physically prepare soil before planting crops.
Whilst important, excessive tillage is also harmful to soil health as it disrupts the
soil’s physical properties, including fragmenting soil aggregates and breaking apart
the plant roots and fungal mycorrhizal networks that bind together soil aggregates.
Tillage can also reduce the coverage of soils by plant residues and top-cover, thereby
exposing soils to erosion and exposing underground organic matter to oxidation,
which is then rapidly decomposed and lost, and must be subsequently replaced by
input organic matter. For example, farming of maize and wheat in the highlands of
Central Mexico uses mechanical tillage that quickly exhausts the organic content of
the soils. Accordingly, tillage must be performed carefully and sustainably to ensure
soil health [39, 40].

4.2 Chemical Degradation

The deterioration of soil’s chemical properties, including the depletion of nutrients,


salinization due to poor irrigation practices and acidification due to excess nitrogen
fertilizers, can all reduce soil health. The accumulation of pollutants, such as
herbicides and pesticides, can also lead to chemical degradation. Approximately
18% of the Mexican territory (34.04 million ha) is impacted by soil chemical
degradation (Fig. 3a, b) [30, 31].
Soils in the arid and semi-arid regions of Mexico are often affected by saliniza-
tion. Extensive and long-term irrigation of these regions can introduce dissolved
salts such as chlorides, sulphates of calcium, magnesium, sodium and potassium.
When the water is lost through evaporation, these salts remain within the soils.
Agricultural areas that exhibit highly saline soils and drainage problems include the
states of Chiapas, Nuevo León, Oaxaca, Sonora, Sinaloa, San Luis Potosí, Veracruz
and Zacatecas [25, 41]. By changing the chemical composition of soil and water,
salinization contributes to losses in biodiversity and soil fertility.
Agricultural activities can also accelerate the process of soil acidification through
the introduction of nitrogen fertilizers, including ammonium sulphate and urea.
These nitrogen fertilizers are oxidized by microbes within soils, resulting in the
production of strong acids such as nitric and sulphuric acid. The acidification of soils
can be further exacerbated by the uptake and removal of non-acid cations (such as
Ca, Mg and K) by the plant crops [5, 42].
The state of Yucatán, in Southeast Mexico exhibits the greatest percentage (55%)
of soils affected by chemical degradation. Large regions of Yucatán soils have been
impacted by an excessive use of fertilizers, herbicides and pesticides and the
accumulation of other pollutants. Furthermore, these pollutants leak into the water
aquifers that pervade the region, and thereby spread contamination to affect larger
areas of soil [43].
Agricultural Land Degradation in Mexico 311

4.3 Water and Wind Erosion

Erosion results in the removal of soil mass, organic matter and nutrients that would
be otherwise available to support plant growth. Water erosion occurs when raindrops
displace soil particles, and flowing surface water removes soil particles, along with
nutrients and organic matter. In addition to removing soil, water erosion can change
soil properties, such as texture, structure and nutrient retention, and damage infra-
structure and waterways, e.g. [44]. Sites at higher risk of water erosion are those at
steep gradients, intense cropping and regions subject to extreme climatic events.
Whilst soil erosion is a natural process, it can be accelerated by poor land manage-
ment. For example, excessive tillage can remove crop residue cover, leaving soils more
exposed to water erosion. Practices that increase the water infiltration of soil, such as
cover cropping and tile drainage, can reduce run-off and limit the impact of water
erosion. Practices that slow run-off, such as terraces and buffer strips, also promote the
deposition and capture of suspended sediment before it leaves the field [45].
Water erosion is a pressing issue in Mexico, with almost 12% of the territory
(22.73 million ha) impacted by this process (Fig. 3a, b). Water erosion is particularly
important in the state of Guerrero on the Southwest Pacific (32% of its land),
however, the impact of water erosion is also broadly apparent across the Sierra
Madre Occidental, the state of Chihuahua and the state of Durango.
Strong wind blowing across light textured soils can also remove soil particles and
result in wind erosion. Sites that have little vegetation cover, low humidity and are
subject to strong winds are all vulnerable to erosion. Nationally, 18.12 million ha
(9.5%) of Mexican territory is impacted by wind erosion, with arid regions, such as
Chihuahua in the Northern plains of Mexico, particularly affected [30, 31]. An effective
method to mitigate wind erosion is to build and maintain vegetative cover. Trees and
shrubs can reduce wind speed, contribute to biodiversity, provide shelter for livestock
and protect from desiccation from the wind. The maintenance of crop residues and
stubble cover due to reduced tillage can also protect soil against wind erosion [46].
The abandonment of the lands used to previously grow alfalfa, oranges, beans, peas
and corn in the Baja California Peninsula has led to wind erosion problems and even the
formation of active sand dunes. This impact of wind erosion is particularly significant in
the Vizcaino Desert and the region bordering the Pacific Ocean on the Plains of
Magdalena. The ongoing wind erosion in Baja California results in the constant removal
and deposition of sediments and prevents establishment of vegetation coverage and the
accumulation of organic matter, and thereby limiting further soil development [47].

5 Examples of Soil Degradation Across Mexican


Ecoregions

Mexico can be broadly categorized into seven ecoregions according to the climate,
physiography, and assemblages of vegetation and faunal communities. These
ecoregions include: (1) the great plains, (2) the southern semi-arid highlands,
312 N. S. Santini et al.

Fig. 4 Mexico largely encompasses seven ecoregions according to the climate, physiography and
assemblages of vegetation and faunal communities. (a) Ecoregions across Mexican states. (b)
Agricultural and urban areas that occur within each ecoregion for 2007, 2012 and 2015 [30, 32,
48, 49]

(3) the North American deserts, (4) tropical and subtropical dry forests, (5) Mediter-
ranean California, (6) temperate highlands and (7) the tropical and subtropical moist
forests. Here, we describe examples of how agriculture impacts soil health across
iconic Mexican ecoregions (Fig. 4a).
Agricultural Land Degradation in Mexico 313

5.1 Mexican Temperate Highlands

Temperate highland forests cover almost a quarter of the Mexican landscape, over
43 million ha, and are dominated by conifer (e.g. Pinus spp., Abies spp., Juniperus
deppeana) and broadleaf trees (e.g. Quercus castanea, Q. rugosa, Q. crassipes).
These highlands extend from 1,000 m above sea level (masl) up to 3,700 masl, where
the subalpine grasslands start [37, 50]. These forests are distributed within the Sierra
Madre Occidental, Sierra Madre Oriental, the Trans-Mexican Volcanic Belt, the
Sierra Madre del Sur and across the highlands from Chiapas (Fig. 4a).
Mexican temperate highlands are habitat to over 7,000 plant species, including
endemic pine (46) and oak (109) species, and are a priority ecoregion for global
conservation [51]. This ecoregion is the habitat of a large number of iconic animals,
including the hibernation site of the monarch butterfly (Danaus plexippus) [50]. The
temperate highland forests also provide key environmental services to large rural and
urban Mexican populations that live within adjacent valleys, where they are key to
ensuring soil fertility to adjacent agricultural lands, carbon sequestration and water
capture.
The Mexican temperate highland regions retain a range of soils, from Andosols
that result from volcanic ash with low-bulk densities and high organic matter content
to Leptosols that are composed of coarse materials and especially susceptible to
erosion and desiccation. Phaeozems, Cambisols, Regosols and Luvisols are also
common across these highlands [52].
Despite their importance, more than 50% of the original Mexican temperate
forests have been deforested for agricultural use, cattle ranching and human settle-
ments (Fig. 4b). Much of the agricultural land that has been converted from
temperate forests is for maize production which has increased in response to a
growing population, as well as increasing demand for use as poultry feed
[53, 54]. Felled timber from the forests is used for fuel, construction or manufactur-
ing [55, 56]. Other common crops across the temperate highlands include peas,
potatoes, broad-beans and black-beans, whilst in the Mexican southwest, temperate
forests are increasingly being converted to avocado plantations [57, 58].
Deforestation causes profound changes to the properties of soils and is a major
driver of soil degradation. By removing the vegetative cover, deforestation reduces
soil organic carbon inputs, increases nutrient leaching from the soil, whilst the use of
heavy machinery reduces soil porosity [59]. The felling of trees also exposes topsoil
to erosion by wind and water by removing windbreaks or channelling streams.
Deforestation can also result in greater soil acidity, with net nutrient export by
harvested products and leaching losses causing a net release of hydrogen ions that
reduce soil pH over time [59].
The temperate highlands of Michoacán have fertile soils that include Andosols
(64% of the area), Luvisols (19.3%), Cambisols (8.3%), Phaeozems (5.3%) and
Vertisols (2.4%) [60]. However, much of these temperate forests have been
converted to agricultural land to produce avocados. The state of Michoacán has
become a major exporter and produces 80% of the avocadoes consumed globally,
314 N. S. Santini et al.

generating US$2.5 billion a year [61]. However, this has been a primary cause of the
loss of primary temperate forests in Michoacán and has resulted in widespread soil
degradation, over-extraction of the water reservoirs and loss of biodiversity [57, 62].
The deforestation of Michoacán’s temperate forests and the conversion of over
234,000 ha of land to avocado plantations have widely impacted soil health. A study
within the Pátzcuaro basin in Michoacán state found that soil erosion increased from
0 to 1.5 tonnes per hectare every year to 3–25 tonnes per hectare every year due to
deforestation [62, 63]. Soil erosion in avocado farms is often most pronounced at the
beginning of the plantation, when the soils are bare. Due to the steep slopes of
temperate highlands, soils are carried off-site, causing deep gullies or channels, and
carrying pesticides that are eventually deposited in water bodies, such as the
Zirahuén Lake in Michoacán (Fig. 1e) [64].
The deforestation of the Michoacán temperate highlands and conversion to
avocado plantations have also reduced the nutrients retained within the soil. Soils
of temperate forest store more organic carbon than those converted to avocado
plantations, which accumulate higher concentrations of nitrogen due to fertilizers.
In temperate forests, there is a higher concentration of nitrogen at lower depths,
whilst in avocado plots, nitrogen concentration is highest in the superficial soil. This
is likely because microorganisms involved in nitrate transformation are obligate
aerobes, and therefore require soil aeration [57].
Avocado plantations have also resulted in the widespread use of herbicides and
pesticides that accumulate within the soils. For example, aluminium phosphide is
commonly applied to control the insect pest Dendroctonus spp. and has been found
to accumulate and pollute aquifers in the area. Adequate soil management practices
are required to promote the conservation of the Andosols capacity to buffer the
lixiviation of agrochemicals, and thereby decreasing the contamination of the
aquifers [65].

5.2 Tropical and Subtropical Dry Forests

Tropical and subtropical dry forests largely occur at sea level and are distributed
across the states of Campeche, Yucatán, Quintana Roo, San Luis Potosi, Veracruz,
Hidalgo and in the Pacific Coast of Mexico from Sinaloa to Chiapas, covering 13%
(almost 32 million ha) of the Mexican landscape (Fig. 4a). This ecoregion is
characterized by a high number of endemic species, including those from the
genus Bursera spp., and a range of cacti species. By 2015, 7.5 million ha of primary
forests was already converted to agricultural land, whilst human/urban settlements
increased from 208,515 ha in 2007 to 291,143 ha in 2015 (Fig. 4b).
Soils within these tropical and subtropical dry forests are often poorly developed
and originate from calcareous-, metamorphic- and volcanic rock. These soils often
have a shallow profile (<25 cm) and can be clayey or sandy. However, much of the
soil across this region exhibits physical and chemical degradation, largely due to
poor land management practices [50]. These forests are often initially felled and
Agricultural Land Degradation in Mexico 315

converted to pasture and agricultural land using slash-and-burn clearing practices in


the hills and heavy machinery in the plains. The subsequent abandonment of
unproductive lands after a few years results in a mosaic landscape between thorny
plant communities and secondary forests that persist in the long-term (Fig. 1f,
g) [66].
The Yucatán Peninsula has a unique karstic geology and is covered by shallow
soils on limestone with an underlying groundwater network [34]. Despite these
shallow, poorly developed soils, and close underlying rock, there is extensive
agriculture production and pasture on the peninsula, including maize, sorghum,
pumpkins and bean farming. However, due to the limited, shallow soils, soil
degradation is a constant concern.
The tropical dry forest of the Yucatán Peninsula is also subject to some of the
highest rates of deforestation across the country [7]. Starting in 2001, much of this
deforestation converted forests to grasslands for large-scale livestock and commer-
cial maize production that was driven, in part, by government agricultural subsidy
programs, migration and population growth. Fires, often used during slash-and-burn
clearing, are a major cause of deforestation across the Yucatán Peninsula, particu-
larly in the northern and central regions of Quintana Roo. These bushfires in this
region often oxidize almost all organic matter and even cause the breakdown of
minerals in topsoils.
In 2012, deforestation across this ecoregion was exacerbated by the expansion of
transgenic soybean farming that largely relies on the herbicide glyphosate, fertilizers
and other organochlorine pesticides [27]. Urban and infrastructural growth for
tourism, such as along the coast of the Riviera Maya and in Cancún, have also
contributed towards ongoing deforestation and subsequent soil degradation.
From May to October tropical storms are common across the Yucatán Peninsula,
causing flooding of the large agricultural regions. This run-off, which often contains
dissolved fertilizers and herbicides, can pollute soils and karstic aquifers. To mitigate
flooding, farmers construct infiltration wells that collect rain water run-off from
fields [67]. Both glyphosate and aminomethylphosphonic acid (a glyphosate prod-
uct) have accumulated within soils and the aquifers across the peninsula, resulting in
long-term soil pollution [67, 68]. The excessive use of glyphosate in the Yucatán
Peninsula has also caused large-scale bee mortality, reducing beekeeping and honey
production that are an important activity and source of income for the local Mayan
communities [27].

5.3 Degradation of North American Deserts

North American deserts extend over a third of the Mexican territory within an area of
55.7 million ha. This ecoregion occurs across the states of Coahuila, Baja California
Sur, Sonora, Sinaloa, Chihuahua, Zacatecas, Durango, San Luis Potosi, Nuevo Leon
and Tamaulipas at lowland elevations. This ecoregion is characterized by dry and
316 N. S. Santini et al.

arid plains and deserts with mean annual rainfall ranging from 130 to 380 mm and
temperatures that can reach over 40°C during the summer months.
The soils across North American deserts largely comprise Regosols that are
weakly developed mineral soils. Accordingly, these soils are fragile and susceptible
to erosion and salinization. Vegetative cover can mitigate the impact of wind and
water erosion. However, the removal of vegetative cover through clearing or over-
grazing can be difficult to address, with plant succession slower in desert ecosystems
than in wet tropical ecosystems. Despite their shallow and poorly developed soils,
North American deserts have been heavily exploited for intensive agricultural
practices, with over 8% of this ecoregion converted for agricultural use (Fig. 4a,
b) [23].
An example of intensive agriculture within this ecoregion occurs within the
Cuatro Ciénegas Valley, within the state of Coahuila. Cuatro Ciénegas is a Ramsar
wetland of international importance and was declared a Flora and Fauna Conserva-
tion Area in 1997 to conserve 84,347 ha that contains a high number of endemic
cacti, reptiles, invertebrates and fish species, as well as to preserve some of the oldest
stromatolites on Earth [69, 70]. Despite the region receiving low precipitation at
200 mm per year, the wetland is supported by a large network of groundwater
streams and lakes.
Farming in the Cuatro Ciénegas Valley involves the intensive production of cattle
feeding crops, with more than half of agricultural land assigned to the production of
alfalfa (Medicago sativa), with the remaining land used for potatoes, maize, wheat
and grape. However, alfalfa production in these arid regions demands large amounts
of water, and an estimated 12 million m3 of water is required to irrigate the alfalfa
plantations each year [71]. This water depletes natural water resources in the region
that would otherwise support the natural vegetation of Cuatro Ciénegas (Fig. 1h, i).
The continuous irrigation of the regions has also increased the leaching of salt
through the soil profile and driven salination of soils in Cuatro Ciénegas. Soil
salination changes the acidity of soils which affects the availability of nutrients to
support plants. For example, critical nitrogen and phosphorus are only available at
pH of 6.5–7 [72]. Comparisons of native desert grassland to cultivated land used for
alfalfa production showed the impact of salination, with cultivated plot having a
lower soil pH. The intensive use of nitrogen amendments promotes nitrification by
releasing hydrogen ions into the soil solution. Together, these changes in the
chemical composition have resulted in a marked deterioration of soil health in Cuatro
Ciénegas [24].

6 Sustainable Agriculture

The Mexican population is expected to grow to 138 million people by 2050 [73]. To
meet the requirements of this growing population, food supplies and agricultural
productivity must increase, and distribution and access to food improve. Healthy
soils will be needed to support these increases in agricultural productivity.
Agricultural Land Degradation in Mexico 317

Accordingly, this growth in agriculture must be achieved using sustainable practices.


Sustainable practices are also needed to reduce greenhouse gas emissions, to reduce
biodiversity and habitat loss, to prevent unsustainable water use and to eliminate
pollution from agricultural chemicals.
Achieving sustainable agriculture is difficult and complex. Sustainable practices
must balance multiple interests, including economic, social and environmental
considerations. Nevertheless, intensive farming practices that damage soils are not
sustainable, as these degrade the finite soil resource and ultimately limit further
farming. Often the short-term benefits from intensive soil use result in soil degrada-
tion, with the long-term loss of soil health and environmental services. Accordingly,
sustainable soil management practices that prevent soil deterioration in the first place
are typically the most cost-effective strategy [74, 75].
Several sustainable agricultural practices can be used to sustain soil health in
Mexico. Traditional farming methods have previously removed crop residue by
burning. However, the burning of residues releases large carbon emissions and
removes organic carbon and nutrients from the soil. As a sustainable alternative,
mulching leaves crop residues on the topsoil surface after harvesting. This residue
protects the soil from water and wind erosion and provides structure for better water
filtration and retention [39, 40].
Reduced and minimal tillage has also been shown to improve soil health.
Reduced tillage prevents the exposure of soil organic content, which is otherwise
oxidized and lost. Excessive tillage prevents earthworm proliferation and organic
carbon accumulation in topsoils. In contrast, reduced tillage maintains the soil
structure, including the networks of fungi and soil biota that retain soil aggregation,
thereby preventing water and wind erosion [39].
Many studies have explored the capacity of conservation tillage systems to
remediate eroded soils. Continuous no-till cropping systems with cover crops have
been found to be particularly effective because of their ability to quickly enhance
levels of organic matter near the surface. Elevated organic matter levels in the
superficial layers of eroded soil can dramatically increase water infiltration, nutrient
cycling and resistance to further erosion. For example, [39] found that degraded soils
in the Patzcuaro watershed in Michoacán improved after 2 years of retention of crop
residues and no-tillage. These improvements included increased organic soil carbon,
microbe biodiversity and water retention. Similarly, [40] also found that no-tillage
and crop residue improved soil aggregates and soil moisture in temperate highlands
of Central Mexico where maize and wheat are common crops.
Crop rotation is also a sustainable practice that can improve soil health. Diverse
and varied crops reduce the establishment and impact of pests, can return organic
matter and fix nutrients within the soil. For example, the rotation of legumes to fix
nitrogen in soils has been a traditional farming practice since pre-Hispanic times.
Traditional farming practices also include the grazing by livestock at sustainable
rates, which can help control weeds and recreate traditional species-rich grasslands,
in which increased plant diversity enhances forage production [16]. These traditional
approaches also replace inorganic fertilizers with farmyard manure that improve
organic matter in soils.
318 N. S. Santini et al.

Ecological restoration aims to recover ecosystem characteristics that have been


degraded or destroyed due to human activity. The use of ecological restoration in
agriculture, termed agroecology, aims to mimic natural ecosystems by increasing
crop biodiversity, fostering natural pollination and pest-control measures and restor-
ing sites to natural vegetation, e.g. [74, 76]. These practices will be needed to reverse
the widespread damage to soil and environmental health caused by expansive
agricultural practices over the past decades across Mexico.

7 Conclusions

Agriculture is a major component of the Mexican economy and has expanded


substantially over the past decades. To continue this growth and increase agricultural
productivity, the Mexican government outlined the ‘National Agricultural Plan for
the period 2017–2030’. The plan aims to improve access of Mexican farmers to
international markets and increase the intensification of agricultural practices. The
plan identified 38 strategic crops, which currently represent approximately 70% of
the national harvested area, and identified potential new regions where these agri-
cultural lands for these strategic crops could expand. However, this expansion and
intensification of agriculture that is described in the plan is likely to increase soil
degradation, ultimately resulting in losses in biodiversity, ecosystem functions and
services that support humans.
The expansion of agriculture into sensitive ecosystems in Mexico over the past
decades has had detrimental effects on biodiversity, carbon storage and important
environmental services. This impact has been well-documented, and it is imperative
that future agricultural programs, such as the ‘National Agricultural Plan
2017–2030’, are managed more sustainably. This sustainable management requires
informed and coordinated research, policy and leadership from Mexican institutions,
including universities, forestry departments, government agencies and local com-
munities to promote sustainable practices.
To promote the sustainable management of the Mexican ecosystems and soils, the
Mexican government has designed the ‘Program for the Sustainable Rural Devel-
opment 2020–2024’. This program aims to reverse land degradation and to incen-
tivize sustainable agricultural practices among small-scale farmers or ‘campesinos’.
These practices include crop rotation, conservation of local genetic resources and
forestry practices that utilize and promote the planting of endemic species, aiming at
maintaining biodiversity and natural landscapes and healthy soils. Despite the
importance of these recent efforts, unsustainable agricultural farming may continue
across Mexican ecoregions unless new policy regulations, such as supervising
chemical and fertilizer use, are urgently applied, most importantly among intensive
agricultural land.
The application of sustainable management of soil practices must be nested
within local communities. Notable examples of conservation and sustainable
resource-use programs have been established by local communities from across
Agricultural Land Degradation in Mexico 319

the State of Mexico, Chiapas, Nayarit and Tabasco that have joined the voluntary
carbon market for forest conservation through the Mexican Carbon Platform (MEX-
ICO2). These examples of local governance and sustainable practices are essential if
we are to conserve the health of Mexican soils.
The complex, mountainous landscape of Mexico harbours many different
ecoregions. As a result, Mexico is one of the most biodiverse countries on Earth
and is the centre of origin of many major global crops. Mexico has also been a cradle
for agriculture, which has been supported by a diverse range of soils that have
developed from the diverse landscapes across the country. However, decades of
deforestation and intensified clearing for agricultural land have indelibly altered the
landscape and impacted soil health. To continue to support agriculture, and more
broadly the diversity and health of the unique Mexican ecosystems, it is imperative
that practices to conserve soil health are urgently established.

References

1. Amundson R, Berhe AA, Hopmans JW, Olson C, Sztein AE, Sparks DL (2015) Soil and human
security in the 21st century. Science 348:647–653. https://doi.org/10.1126/science.1261071
2. Rabot E, Wiesmeier M, Schlüter S, Vogel HJ (2018) Soil structure as an indicator of soil
functions: a review. Geoderma 314:122–137. https://doi.org/10.1016/j.geoderma.2017.11.009
3. Totsche KU, Amelung W, Gerzabek MH, Guggenberger G, Klumpp E, Knief C, Lehndorff E,
Mikutta R, Peth S, Prechtel A, Ray N, Kögel-Knabner I (2018) Microaggregates in soils. J Plant
Nutr Soil Sci 181:104–136. https://doi.org/10.1002/jpln.201600451
4. Chen S, Ai X, Dong T, Li B, Luo R, Ai Y, Chen Z, Li C (2016) The physico-chemical
characteristic of artificial soil for cut slope restoration in Southwestern China. Sci Rep 6:
20565. https://doi.org/10.1038/srep20565
5. Weil RR, Brady NC (2016) The nature and properties of soils.15th edn. Pearson
6. Trap J, Bonkowski M, Plassard C, Villenave C, Blanchart E (2016) Ecological importance of
soil bacterivores for ecosystem functions. Plant Soil 398:1–24. https://doi.org/10.1007/s11104-
015-2671-6
7. Comisión Nacional Forestal (2020) Estimación de la tasa de deforestación en México para el
periodo 2001–2018 mediante el método de muestreo. Documento Técnico, Jalisco
8. Nannipieri P, Ascher J, Ceccherini MT, Landi L, Pietramellara G, Renella G (2017) Microbial
diversity and soil functions. Eur J Soil Sci 68:1–26. https://doi.org/10.1111/ejss.4_12398
9. Obalum SE, Chibuike GU, Ouyang Y (2017) Soil organic matter as sole indicator of soil
degradation. Environ Monit Assess 189:176. https://doi.org/10.1007/s10661-017-5881-y
10. Lal R (2015) Restoring soil quality to mitigate soil degradation. Sustain 7:5875–5895. https://
doi.org/10.3390/su7055875
11. Stavi I, Lal R (2015) Achieving zero net land degradation: challenges and opportunities. J Arid
Environ 112:44–51. https://doi.org/10.1016/j.jaridenv.2014.01.016
12. Secretaría de Agricultura y Desarrollo Rural (2017) Planeación Agrícola Nacional 2017–2030.
Secretaría de Agricultura y Desarrollo Rural, Mexico. https://www.gob.mx/agricultura/
acciones-y-programas/planeacion-agricola-nacional-2017-2030-126813
13. Bellon MR, Mastretta-Yanes A, Ponce-Mendoza A, Ortiz-Santamaría D, Oliveros-Galindo O,
Perales H, Acevedo F, Sarukhán J (2019) Evolutionary and food supply implications of ongoing
maize domestication by Mexican campesinos. Proc R Soc B Biol Sci 285:20181049. https://doi.
org/10.1098/rspb.2018.1049
320 N. S. Santini et al.

14. Servicio de Información Alimentaria y Pesquera (2019) Producción agrícola. Servicio de


Información Alimentaria y Pesquera, Mexico. https://www.gob.mx/siap
15. Ebel R, Pozas Cárdenas JG, Soria Miranda F, Cruz González J (2017) Manejo orgánico de la
milpa: rendimientos de maíz, frijol y calabaza en monocultivo y policultivo. Terra Latinoam 35:
149–160
16. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (2019) Alimentar a México
sin deforestar. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, México
17. Arteaga MC, Moreno-Letelier A, Mastretta-Yanes A, Vázquez-Lobo A, Breña-Ochoa A,
Moreno-Estrada A, Eguiarte LE, Piñero D (2016) Genomic variation in recently collected
maize landraces from Mexico. Genomics Data 7:38–45. https://doi.org/10.1016/j.gdata.2015.
11.002
18. Bocco G, Castillo BS, Orozco-Ramírez Q, Ortega-Iturriaga A (2019) La agricultura en terrazas
en la adaptación a la variabilidad climática en la Mixteca Alta, Oaxaca, México. J Lat Am Geogr
18:141–168. https://doi.org/10.1353/lag.2019.0006
19. Pantoja A, Smith-Pardo A, García A, Sáenz A, Rojas F (2014) Principios y avances sobre
polinización como servicio ambiental para la agricultura sostenible en países de Latinoamérica
y El Caribe. Organización de las Naciones Unidas para la Alimentación y la Agricultura, Chile
20. Aguilar JC, Tolón BA, Lastra BX (2011) Evaluación integrada de la sostenibilidad ambiental,
económica y social del cultivo de maíz en Chiapas, México. Rev la Fac Ciencias Agrar 43:155–
174. http://www.redalyc.org/articulo.oa?id=382837648011
21. Orozco-Ramírez Q, Astier M, Barrasa S (2017) Agricultural land use change after NAFTA in
Central West Mexico. Land. https://doi.org/10.3390/land6040066
22. Peel DS, Mathews KH, Johnson RJ (2012) Trade, the expanding Mexican beef industry, and
feedlot and stocker cattle production in Mexico, a report from USDA – Economic Research
Service
23. Bonilla-Moheno M, Aide TM (2020) Beyond deforestation: land cover transitions in Mexico.
Agric Syst 178:102734. https://doi.org/10.1016/j.agsy.2019.102734
24. Tapia-Torres Y, Ortiz PC, Hernández-Becerra N, Morón Cruz A, Beltrán O, García-Oliva F
(2018) How do agricultural practices modify soil nutrient dynamics in CCB? In: García-
Oliva F, Elser J, Souza V (eds) Ecosystem ecology and geochemistry of Cuatro Cienegas.
Cuatro Ciénegas basin: an endangered hyperdiverse oasis. Springer, Cham. https://doi.org/10.
1007/978-3-319-95855-2_12
25. Endo T, Yamamoto S, Larrinaga JA, Fujiyama H, Honna T (2011) Status and causes of soil
salinization of irrigated agricultural lands in Southern Baja California, Mexico. Appl Environ
Soil Sci. https://doi.org/10.1155/2011/873625
26. Hernández-Becerra N, Tapia-Torres Y, Beltrán-Paz O, Blaz J, Souza V, García-Oliva F (2016)
Agricultural land-use change in a Mexican oligotrophic desert depletes ecosystem stability.
PeerJ 4:e2365. https://peerj.com/articles/2365/
27. Villanueva-Gutiérrez R, Echazarreta-González C, Roubik DW, Moguel-Ordóñez YB (2014)
Transgenic soybean pollen (Glycine max L.) in honey from the Yucatán peninsula, Mexico. Sci
Rep 4:4022. https://doi.org/10.1038/srep04022
28. Mendivil-Garcia K, Amabilis-Sosa LE, Rodríguez-Mata AE, Rangel-Peraza JG, González-
Huitron V, Cedillo-Herrera CIG (2020) Assessment of intensive agriculture on water quality
in the Culiacan River basin, Sinaloa, Mexico. Environ Sci Pollut Res 27:28636–28648. https://
doi.org/10.1007/s11356-020-08653-z
29. Krasilnikov P, del Carmen Gutiérrez-Castorena M, Cruz-Gaistardo CO et al (2013) Soils of
Mexico. Springer
30. Instituto Nacional de Estadística y Geografía (2007) Conjunto de datos vectoriales. Scale 1:
250,000. Continuo Nacional, México. https://www.inegi.org.mx/temas/edafologia/#Mapa
31. Secretaría de Medio Ambiente y Recursos Naturales (2004) Degradación del suelo en la
República Mexicana. Scale: 1:250 000, Mexico. http://geoportal.conabio.gob.mx/metadatos/
doc/html/degra250kgw.html
Agricultural Land Degradation in Mexico 321

32. Instituto Nacional de Estadística y Geografía (2021) Áreas geoestadísticas estatales. Scale 1:
250000.1st edn, Aguascalientes. https://www.inegi.org.mx/temas/mg/
33. Astier-Calderón M, Maass-Moreno M, Etchevers-Barra J (2002) Derivación de indicadores de
calidad de suelos en el contexto de la agricultural sustentable. Agrociencia 366:605–620. http://
www.redalyc.org/articulo.oa?id=30236511
34. Perry E, Velazquez-Oliman G, Marin L (2002) The hydrogeochemistry of the karst aquifer
system of the Northern Yucatán Peninsula, Mexico. Int Geol Rev 44:191–221. https://doi.org/
10.2747/0020-6814.44.3.191
35. Siebe C, Bocco G, Sánchez J, Velázquez A (2003) Suelos: distribución, características y
potencial de uso. In: Velázquez A, Torres A, Bocco G (eds) Las enseñanzas de San Juan:
investigación participativa para el manejo integral de recursos naturales. Secretaría de Medio
Ambiente, Recursos Naturales, Instituto Nacional de Ecología, México
36. Torres Guerrero CA, Castorena M, Ortiz Solorio CA, Gutiérrez Castorena EV (2016) Agricul-
tural management of Vertisols in Mexico: a review. Terra Latinoam 34:457–466
37. Santini NS, Adame MF, Nolan RH, Miquelajauregui Y, Piñero D, Mastretta-Yanes A, Cuervo-
Robayo AP, Eamus D (2019) Storage of organic carbon in the soils of Mexican temperate
forests. For Ecol Manag 446:115–125. https://doi.org/10.1016/j.foreco.2019.05.029
38. Mendoza E, Fay J, Dirzo R (2005) A quantitative analysis of forest fragmentation in Los
Tuxtlas, Southeast Mexico: patterns and implications for conservation. Rev Chil Hist Nat 78:
159–187. https://www.redalyc.org/articulo.oa?id=369944275008
39. Roldán A, Caravaca F, Hernández MT, García C, Sánchez-Brito C, Velásquez M, Tiscareño M
(2003) No-tillage, crop residue additions, and legume cover cropping effects on soil quality
characteristics under maize in Patzcuaro watershed (Mexico). Soil Tillage Res 72:65–73.
https://doi.org/10.1016/S0167-1987(03)00051-5
40. Govaerts B, Sayre KD, Goudeseune B, De Corte P, Lichter K, Dendooven L, Deckers J (2009)
Conservation agriculture as a sustainable option for the central Mexican highlands. Soil Tillage
Res 103:222–230. https://doi.org/10.1016/j.still.2008.05.018
41. González-Acevedo ZI, Padilla-Reyes DA, Ramos-Leal JA (2016) Quality assessment of irri-
gation water related to soil salinization in Tierra Nueva, San Luis Potosí, Mexico. Rev Mex de
Cienc Geolo 33:271–285
42. Kunhikrishnan A, Thangarajan R, Bolan NS, Xu Y, Gleeson DB, Seshadri B, Zaman M,
Barton L, Tang C, Luo J, Dalal R, Ding W, Kirkham MB, Naidu R (2016) Functional
relationships of soil acidification, liming, and greenhouse gas flux. In: Advances in agronomy,
vol 139. https://doi.org/10.1016/bs.agron.2016.05.001
43. Hernández-Terrones L, Rebolledo-Vieyra M, Merino-Ibarra M (2011) Groundwater pollution
in a Karstic region (NE Yucatán): baseline nutrient content and flux to coastal ecosystems.
Water Air Soil Pollut 218:517–528. https://doi.org/10.1007/s11270-010-0664-
44. Xiong M, Sun R, Chen L (2018) Effects of soil conservation techniques on water erosion: a
global analysis. Sci Total Environ 645:753–760. https://doi.org/10.1016/j.scitotenv.2018.
07.124
45. Cotler H, Martínez M, Etchevers JD (2016) Carbono orgánico en suelos agrícolas de México:
Investigación políticas públicas. Terra Latinoam 34:125–138
46. Seitz S, Goebes P, Puerta VL, Pujol Pereira EI, Wittwer R, Six J, van der Heijden MGA,
Scholten T (2019) Conservation tillage and organic farming reduce soil erosion. Agron Sustain
Dev 39:4. https://doi.org/10.1007/s13593-018-0545-z
47. Pretelín VI, Bearden K, Bird JL (2021) Evaluación del sistema agrícola y alimentario de BCS.
Catalizar una región alimentaria local próspera. Alianza para la Seguridad Alimentario de Baja
California Sur A.C. Baja California Sur
48. Instituto Nacional de Estadística y Geografía (2013) Conjunto de datos vectoriales de uso de
suelo y vegetación. Scale 1:250 000, series V. Capa unión, Aguascalientes. https://www.inegi.
org.mx/temas/usosuelo/#Descargas
322 N. S. Santini et al.

49. Instituto Nacional de Estadística y Geografía (2016) Conjunto de datos vectoriales de uso de
suelo y vegetación. Scale 1:250 000, series VI. Capa unión, Aguascalientes. https://www.inegi.
org.mx/temas/usosuelo/#Descargas
50. Challenger A, Soberon J (2008) Los ecosistemas terrestres. In Capital natural de México, vol
I. Conocimiento actual de la biodiversidad. Comisión Nacional para el Conocimiento y Uso de
la Biodiversidad, México
51. Valencia SA (2004) Diversidad del género Quercus (Fagaceae) en México. Bol Soc Bot Méx
75:33–53. http://redalyc.org/articulo.oa?id=57707503
52. Comisión Nacional Forestal (2018) Inventario Nacional Forestal y de Suelos (2009–2014).
https://snigf.cnf.gob.mx/resultados-2009-2014-resultados-que-recaba-los-principales-
indicadores-forestales-generados-a-partir-del-analisis-estadistico-de-las-variables-levantadas-
en-campo/
53. Galicia L, García-Romero A (2007) Land use and land cover change in Highland temperate
forests in the Itzta-Popo National Park, Central Mexico. Mount Res Dev 27:48–57. https://doi.
org/10.1659/0276-4741(2007)27[48:LUALCC]2.0.CO;2
54. Boué C, López Ridaura S, Rodríguez Sánchez LM, Hellin J, Fuentes Ponce M (2018) Local
dynamics of native maize value chains in a peri-urban zone in Mexico: the case of San Juan
Atzacualoya in the state of Mexico. J Rural Stud 64:28–38. https://doi.org/10.1016/j.jrurstud.
2018.09.014
55. Sarukhán J, Koleff P, Carabias J et al (2017) Capital Natural de México. Síntesis: evaluación del
conocimiento y tendencias de cambio, perpsectivas de sustentabilidad, capacidades humanas e
institucionales. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, México
56. Santini NS, Villarruel-Arroyo A, Adame MF, Lovelock CE, Nolan RH, Gálvez-Reyes N,
González EJ, Olivares-Resendiz B, Mastretta-Yanes A, Piñero D (2020) Organic carbon stocks
of Mexican montane habitats: variation among vegetation types and land-use. Front Environ Sci
8. https://doi.org/10.3389/fenvs.2020.581476
57. Bravo-Espinosa M, Mendoza ME, Carlón Allende T, Medina L, Sáenz-Reyes T, Páez R (2014)
Effects of converting forest to avocado orchards on topsoil properties in the Trans-Mexican
volcanic system, Mexico. L Degrad Dev 25:452–467. https://doi.org/10.1002/ldr.2163
58. Hernández-Cumplido J, Rodriguez-Saona C, Ruíz-Rodríguez CE, Guevara-Fefer P, Aguirre-
Paleo S, Miranda Trejo S, Callejas-Chavero A (2021) Genotypic variation in plant traits,
chemical defenses, and resistance against insect herbivores in avocado (Persea americana)
across a domestication gradient. Front Agron 2:1–12. https://www.frontiersin.org/arti
cle/10.3389/fagro.2020.616553
59. Veldkamp E, Schmidt M, Powers JS, Corre MD (2020) Deforestation and reforestation impacts
on soils in the tropics. Nat Rev Earth Environ 1:590–605. https://doi.org/10.1038/s43017-020-
0091-5
60. Dubrovina IA, Bautista F (2014) Analysis of the suitability of various soil groups and types of
climate for avocado growing in the state of Michoacán, Mexico. Eurasian Soil Sci 47:491–503.
https://doi.org/10.1134/S1064229314010037
61. Statista (2021) Production of avocado in Mexico from 2010 to 2020. https://www.statista.com/
statistics/591329/mexico-fresh-avocado-production/
62. Chávez-León G, Tapia Vargas LM, Bravo Espinoza M et al (2012) Impacto del cambio de uso
del suelo forestal a huertos de aguacate. Instituto Nacional de Investigaciones Forestales,
Agrícolas y Pecuarias. INIFAP, México
63. Gómez-Tagle A, Morales-Chávez R, García-González Y, Gómez-Tagle AF (2019) Partición de
la precipitación en cultivo de aguacate y bosque de pino-encino en Michoacán, México.
Biológicas 21:1–18
64. Chacón A, Rosas C, Rendón M, Cruz O (2010) Balance hidrológico de la cuenca del lago de
Zirahuén. In: Estrada B, Ayala G (eds) Espejo de los dioses: estudios sobre ambiente y
desarrollo en la cuenca del Lago de Zirahuén. Universidad Michoacana de San Nicolás de
Hidalgo, Instituto de Investigaciones Económicas y Empresariales, México
Agricultural Land Degradation in Mexico 323

65. Raymundo E, Nikolskii I, Duwig C, Prado Pano BL, Hidalgo Moreno CI, Gavi Reyes F,
Figueroa Sandoval B (2009) Transporte de atrazina en un Andosol y un Vertisol de México.
Interciencia 34:330–337
66. Trejo I, Dirzo R (2000) Deforestation of seasonally dry tropical forest: a national and local
analysis in Mexico. Biol Conserv 94:133–142. https://doi.org/10.1016/S0006-3207(99)
00188-3
67. Polanco Rodríguez AG, Riba López MI, DelValls CA, Araujo León A, Banik SD (2018) Impact
of pesticides in karst groundwater. Review of recent trends in Yucatán, Mexico. Groundw
Sustain Dev 7:20–29. https://doi.org/10.1016/j.gsd.2018.02.003
68. Ding C, Wang X, Liu H, Li Y, Sun Y, Lin Y, Sun W, Zhu X, Dai Y, Luo C (2018) Glyphosate
removal from water by functional three-dimensional graphene aerogels. Environ Chem 15:325–
335. https://doi.org/10.1071/EN18087
69. Balleza J, Villaseñor JL (2011) Contribución del estado de Zacatecas (México) a la
conservación de la riqueza florística del Desierto Chihuahuense. Acta Bot Mex 89:61. https://
doi.org/10.21829/abm94.2011.271
70. Souza V, Siefert JL, Escalante AE, Elser JJ, Eguiarte LE (2012) The Cuatro Ciénegas Basin in
Coahuila, Mexico: An astrobiological precambrian park. Astrobiology 12:641–647. https://doi.
org/10.1089/ast.2011.0675
71. Mamer E, Newton TB (2017) The relationship between the Cuatrociénegas gypsum dune field
and the regional hydrogeology Coahuila Mexico. New Mexico Bureau of Geology and Mineral
Resources
72. Plaster JE (2005) La ciencia del suelo y su manejo. Internacional Thompson Editores, Madrid
73. Comisión Nacional de Vivienda (2021) https://sniiv.conavi.gob.mx/demanda/poblacion_
proyecciones.aspx
74. Guzmán Luna A, Ferguson BG, Giraldo O, Schmook B, Aldasoro Maya EM (2019) Agroecol-
ogy and restoration ecology: fertile ground for Mexican peasant territoriality? Agroecol Sustain
Food Syst 43:1174–1200. https://doi.org/10.1080/21683565.2019.1624284
75. Cotler H, Corona JA, Galena-Pizaña JM (2020) Erosión de suelos y carencia alimentaria en
México una primera aproximación. Investigaciones Geográficas. https://doi.org/10.14350/rig.
59976
76. Sarandón SJ, Flores CC (2014) Agroecología: Bases Teóricas para el Diseño y Manejo de
Agroecosistemas Sustentables. Editorial de la Universidad de La Plata, Buenos Aires
Agricultural Land Degradation in
South Africa

C. W. van Huyssteen and C. C. du Preez

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
2 Natural Resources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
2.1 Climate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
2.2 Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
2.3 Vegetation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
3 Land Use and Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334
4 Soil Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
4.1 Plinthite Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
4.2 Surface Crusting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
4.3 Subsoil Compaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
4.4 Structural Decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
4.5 Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
4.6 Fertility Decline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
4.7 Elemental Imbalance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
4.8 Organic Matter Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
4.9 Organism Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
4.10 Pathogen Increase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350
5 Land Degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350
6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355

Abstract South Africa receives less than half of the annual global average precip-
itation, resulting in range-, crop-, and forestlands being very prone to degradation.
This dry climate and a very old and stable landscape formed unique combinations of
soil and vegetation. The major land use is therefore extensive grazing (83%), with
only 14% arable land, and 1% forestry. About 2% of this land is severely affected by
wind erosion, whilst water erosion varies from <1 to 60 Mg ha-1 year-1. Natural
acidification occurs on 16 × 106 ha, with anthropogenic acidification on

C. W. van Huyssteen (✉) and C. C. du Preez


University of the Free State, Bloemfontein, South Africa
e-mail: [email protected]; [email protected]

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 325
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 325–362, DOI 10.1007/698_2022_922,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 29 January 2023
326 C. W. van Huyssteen and C. C. du Preez

12.9 × 106 ha cultivated land. About 1.4 × 106 ha irrigated land is saline, whilst
sodicity is not widespread. Water pollution occurs in some rivers in South Africa, but
agriculture is only responsible for 20% of this nitrogen and 53% of the phosphorus,
with the remainder from human effluent and industry. Only 4% of soils contain >4%
organic carbon, whilst 58% contain <0.5% organic carbon. Soil degradation is of
greatest concern in the KwaZulu-Natal, Northern, and Eastern Cape Provinces,
whilst vegetation degradation is of greatest concern in the Northern, KwaZulu-
Natal, and Northern Cape Provinces. Land tenure, inappropriate land use, and
management practices within the unique ecosystems emerged as primary drivers
of agricultural land degradation. This degradation seems to be restricted to certain
geographical areas. Focused management practices for either prevention or improve-
ment can therefore be developed.

Keywords Cropland, Forestland, Rangeland, Soil, Vegetation

1 Introduction

On a global scale, South Africa should be considered an arid country, since it


receives on average less than half (460 mm a-1), against the global average of
990 mm precipitation per annum [1]. General consensus exists amongst experts that
the susceptibility of land (used here to refer to the exposed part of the earth’s surface,
e.g. the atmosphere, climate, rocks, soils, vegetation, animals, and water) to degrade
due to anthropogenic activities is greater under drier than under wetter climates
[2]. Being the remnant of the approximate central part of the Gondwanaland
supercontinent until its breakup about 180 × 106 years before present, Southern
Africa comprises mostly sedimentary shale, sandstone, and mudstone. Following
breakup, the subcontinent underwent several upliftment cycles, resulting in the
major part of South Africa occurring higher than 1,000 masl. This resulted in deep
valleys and steep slopes in the eastern and south-western parts that turn gradually to
a large flat landscape in the central and western parts having patches of alluvial,
colluvial, and aeolian deposits. This set the stage for land quite prone to various
degradation processes [3].
South Africa comprises 122.5 × 106 ha, of which the majority (68%) is used for
extensive ranching, followed by cropping (12%) and forestry (0.8%) [4]. These
farming activities lead in many instances to land degradation, since management
skills and financial capacities vary vastly between practising farmers [1]. For this
discussion, these groups of farmers are defined as follows: Smallholder farmers refer
to those farmers with plot sizes <2 ha in size. Subsistence farmers are farmers who
primarily produce for their own consumption, but may sell some excess produce.
Communal farmers may or may not be smallholder and/or subsistence farmers as
well who do not hold the title deed of the land they farm. Commercial farmers refer
Agricultural Land Degradation in South Africa 327

to farmers who have a profit motive and include single person operations to large
corporate enterprises [5].
The majority of farming enterprises in South Africa are labour rather than capital
intensive. In this regard, a 2017 census [6] of commercial farming units (in this case
registered for value-added tax) showed that only 2,610 of farm enterprises (6.5% of
the total) earned 67% of the agriculture income in South Africa. These farm
enterprises had an annual turnover greater than ZAR22.5 million (about USD1.4
million) each and employed approximately 51% of all employees in the agricultural
sector. About 25,000 farm enterprises (62% of the total commercial enterprises)
earned less than ZAR1 million (about USD60,000) each annually and thus represent
only 2% of the total agricultural income of commercial farming enterprises. In 2018,
commercial farming was practised on 46.4 × 106 ha or 37.9% of the total land area.
This commercial agricultural land comprised 36.5 × 106 ha (82.8%) rangeland,
7.6 × 106 ha (17.2%) cropland, and 2.3 × 106 ha (5.0%) undefined land [6].
Degradation of land (e.g. erosion, compaction, crusting, and organic matter
decline) results in the temporary or permanent reduction of the land’s productive
capacity for farming use (e.g. ranching, cropping, and/or forestry), farming systems
(e.g. smallholder, subsistence, communal, and commercial), and economic value.
Cognisance of land degradation is therefore of critical importance in an effort to
terminate and even to reverse this degradation and to thus conserve this irreplaceable
resource. This will aid in ensuring food security and employment opportunities for
the population of about 58.8 × 106 in South Africa [4].
This chapter starts with a concise overview of the natural resources (climate, soil,
and vegetation) to set the scene for the conditions under which ranching, cropping,
and forestry are practised in South Africa. Then the current land uses and resultant
produce are dealt with, focusing strongly on ranching and cropping, since these are
by far the greatest farming activities. This is followed by a detailed discussion of the
soil degradation processes occurring in range-, crop-, and forestlands. Soil is an
integral part of land and the two are so closely interwoven that the degradation of the
one inevitably implies degradation of the other. Finally, an appraisal of the current
state of land degradation in South Africa, as reflected by soil and vegetation, is
presented.

2 Natural Resources

The use of land for ranching, cropping, and forestry depends largely on the interac-
tion of climate, soil, and vegetation. Only these three natural resources are therefore
concisely discussed here.
328 C. W. van Huyssteen and C. C. du Preez

2.1 Climate

The rainfall that South Africa receives ranges from less than 100 mm a-1 in the
north-western part to over 1,500 mm a-1 in the eastern part of the country (Fig. 1),
averaging 460 mm a-1 against the global average of 990 mm a-1 [1].
Although the annual average rainfall in the mountainous eastern parts exceeds the
global average, the major problem is that rainfall is very poorly distributed through-
out the country and also throughout the year. Additionally, this low rainfall is
exacerbated by high evaporation, following an approximate converse pattern,
increasing from less than 1,800 mm a-1 in the southeast to 3,000 mm a-1 in the
north-west. Winters (June–August) are cold, whilst summers (November–February)
are warm. In winter, the minimum temperatures can drop below freezing, mainly due
to altitude.
The major part of South Africa is characterised by summer (November–March)
rainfall, occurring mainly as intense thundershowers in the afternoon. In the south-
western region of the country, rainfall is predominating in the winter (April–
September) months. Snow is rare throughout South Africa, but may fall on the
mountains of the Southern and Western Cape and in the Drakensberg in the winter

Fig. 1 Mean annual precipitation and evaporation in South Africa, based on data from Schulze
et al. [7]
Agricultural Land Degradation in South Africa 329

months. Hence, the climate of South Africa is generally conducive for the degrada-
tion of agricultural land [1].

2.2 Soil

Soils in South Africa are typically well-formed due to the geologically stable and
extremely old nature of the parent material. As could be expected, a rather good
relationship between soil distribution and parent material as augmented by climate
therefore exists. The general distribution of soils in South Africa is given in Fig. 2
and briefly discussed below. The terminology used here largely follows Fey’s
abridgment [9] of Soil Classification Working Group [10].
Clay soils, including duplex soils, vertic soils, and soils with dark, well-structured
topsoils, occupy the largest part of South Africa. The duplex soils are distinguished
by a distinct increase of clay in the subsoil horizons that results in strong columnar,
prismatic, or blocky structure, typically with an upper horizon boundary that is clear
or abrupt. Duplex soils occur commonly all over South Africa, particularly in the
more arid areas. Vertic soils have swelling, smectite clays, high clay content, and
strong structure and are chemically and physically active. Vertic soils typically occur
on mafic parent material in the semi-arid parts of the country, where the climate is

Fig. 2 Generalised soil distribution in South Africa, based on data from the Land Type Survey
Staff [8]
330 C. W. van Huyssteen and C. C. du Preez

characterised by wet summers and dry winters, allowing for swelling and shrinking
of these soils. The topsoils of melanic soils are strongly structured, have elevated
base status, and have dark colours. Duplex soils are particularly prone to water
erosion since the parent material typically contains sodium associated with the
depositional environment of the shale parent material. This sodium is released
upon weathering of the shale and accumulates in the subsoil, leading to clay
dispersion and erosion, especially when exposed by the removal of the protective
vegetation through ploughing or overgrazing [9, 10].
Red, yellow, and grey plinthic soils are characterised by segregation and accu-
mulation of iron and manganese oxides under alternating water saturated conditions
and may or may not be hardened. These soils therefore typically occur in the wetter
eastern parts, and on almost any parent material. Red and yellow apedal soils may
occur subdominantly in association with the plinthic soils. Sandy variants of these
soils are prone to wind erosion, whilst clayey variants can be prone to water erosion
and acidification [9, 10].
Red and yellow apedal soils are defined based on the homogeneously coloured
red and/or yellow subsoils. They are widely distributed throughout South Africa and
therefore vary greatly in base status due to their occurrence across various climate
zones. Plinthic soils may occur subdominantly in association with these soils. The
humic soils are characterised by accumulation of organic matter under well-drained
conditions and with acidic parent material and therefore occur mainly in the cool,
high rainfall elevated plateaus slightly inland from the coast, and in the mist- covered
midlands of KwaZulu-Natal. The sandy variants of the red, yellow, and grey plinthic
and apedal soils are mostly of aeolian origin on the flat landscapes, resulting in them
being particularly prone to wind erosion, when the protective vegetation is removed
through ploughing or overgrazing, as well as to surface crusting, subsoil compaction,
and acidification. Clayey variants can be prone to water erosion and acidification
[9, 10].
The genesis of soil is a long-term process, hence, little or no soil formation is
observed in the rock and rocky soils, and the original rock or saprolite is therefore
still distinguishable in the lithic soils. Podzolic soils occur in association with the
lithic soils and have illuvial organic material and sesquioxides in the subsoil, without
a marked increase in clay. The distribution of podzolic soils is largely limited to the
high rainfall, mountainous parts of the Western Cape Province, in the southern part
of South Africa, and are therefore naturally acidic. Rock and rocky soils do not form
deep gullies, but the limited topsoil can easily be eroded, leaving the barren rocky
subsoil [9, 10].
Sandy soils incorporate those cumulic soils that form in recent unconsolidated
colluvial or alluvial depositions that have undergone little or no soil formation.
These soils occur widely in the north-western part of South Africa and in association
with rivers and are therefore susceptible to water erosion. However, they are also
prone to wind erosion, especially if the protective vegetation cover is removed
through ploughing or overgrazing, subsoil compaction, and acidification [9, 10].
Swamps and alluvial plains are characterised by gleyic and organic soils, typical
of wetlands, and therefore do not occupy vast areas in South Africa. The gleyic soils
Agricultural Land Degradation in South Africa 331

are characterised by long periods of water saturation that result in the accumulation
or removal of colloidal material: organic soils, with >10% organic carbon that have
accumulated under acidic, wet, and/or cold conditions. Alluvial soils can be prone to
water erosion due to the increased water accumulation in these areas, organic matter
decline, and structural decay [9, 10].
Weakly developed soils include the cumulic, silicic, calcic, and anthropic soils.
Cumulic soils develop in recent unconsolidated alluvial or colluvial deposits that
have undergone some soil formation. These soils occur widely all over South Africa.
Silicic and calcic soils are exclusively found in the arid western parts. The silicic
soils have an indurated subsoil that has silica as the main cementing agent. On the
other hand, calcic soils have accumulated secondary (pedogenic) carbonates of
calcium and/or magnesium, in the subsoil that may or may not be hardened.
Anthropic soils are produced through anthropogenic soil deposition. Weakly devel-
oped soils can be prone to water erosion since these soils typically occur along water
courses where water flow is greater and because some of these soils can have
elevated sodium concentrations which may lead to clay dispersion and erosion,
when the vegetative cover is destroyed [9, 10].

2.3 Vegetation

Vegetation mapping of South Africa started with Pole Evans [11], but it was Acocks
[12] who really laid the foundation for the development of vegetation mapping.
Mucina and Rutherford [13] provide the most recent biome classification and
vegetation map (Fig. 3) of South Africa. They defined and mapped nine biomes,
incorporating Fynbos, Desert, Succulent Karoo, Nama-Karoo, Albany Thicket,
Savanna, Grassland, Forest, and Indian Ocean Coastal Belt.
Finch and Meadows [14] describe the biomes as follows: the Fynbos Biome is
largely restricted to the sandy soils of the Cape Fold Mountains, is endemic to
South Africa, and has ericoid, proteoid, and restioid vegetation. The Desert Biome
occurs as a narrow strip along the Orange River in the extreme north-western part of
South Africa, and has barren areas, dwarf shrubs, and occasional trees, typical of low
rainfall and high temperature. The Succulent Karoo Biome covers the western part of
South Africa, has approximately 5,000 vascular plant species, 40% plant endemism,
and with Aizoaceae and Crassulaceae as the largest plant families. The central
western part of South Africa is dominated by the Nama-Karoo Biome, with grass,
succulent, geophyte, and forb vegetation, which is very susceptible to overgrazing
and resultant soil erosion. The Albany Thicket Biome is located in the southeast of
South Africa, has thick, spiny, semi-succulent bushes and small trees, and is
arguably the most extensively degraded due to biomass reduction, alien plant
invasion, and soil erosion. The Savanna Biome primarily covers the north-western,
northern, and eastern extremities of South Africa, has continuous grass and discon-
tinuous tree cover, and is susceptible to alien plant invasion and bush encroachment.
In the central eastern part of South Africa, the Grassland Biome dominates, with
332 C. W. van Huyssteen and C. C. du Preez

Fig. 3 The biomes of South Africa, based on data adapted from Mucina and Rutherford [13]

grasses dominating, which is threatened by destruction through farming activities


such as ranching, cropping, and forestry. The Forest Biome occurs as small areas on
the escarpment and in the coastal lowlands, comprise Afrotemperate and coastal
forests (with evergreen and semi-deciduous trees), and are subject to destruction and
fragmentation. The Indian Ocean Coastal Belt Biome occupies a thin coastal strip of
eastern South Africa, with extensive grasslands, trees, lianas, and epiphytes, and is
threatened by destruction through the mentioned three farming activities.
As alluded to above, and extensively reviewed by Finch and Meadows [14], each
biome has unique challenges and susceptibility to degradation. The different biomes
also have vastly differing production (Fig. 4), palatability, nutrition, and therefore
different potential stocking rates (Fig. 5). These differences are largely driven by
climate, geology, and soils. Understanding these variables is therefore critical in
managing stocking rates to prevent excessive vegetation removal by livestock [17],
which is the main driver of wind and water erosion in rangelands [18].
The normalised difference vegetation index (NDVI) map (Fig. 4) was compiled
by the ARC-ISCW [15] as a composite of the annual maximum NDVI values (the
month with the greatest growth per year), averaged over 17 years (1986–2003). The
NDVI is used as a measure of the vegetation vigour and biomass production by using
the green bands of satellite images. Averaging these values over the long term
therefore expresses the long-term biomass production potential, imperative for
Agricultural Land Degradation in South Africa 333

Fig. 4 Normalised difference vegetation index (NDVI) map for South Africa, compiled as a
composite of annual maximum values, averaged over 17 years, based on data adapted from
ARC-ISCW [15]

sustaining extensive grazing (see, for example, Fig. 5). The NDVI map is almost a
carbon copy of the rainfall map (Fig. 1), substantiating the vital role of water in
biomass vigour and production. It is this vegetative production that ultimately
supports animal grazing, as expressed by the grazing capacity map (Fig. 5).
The grazing capacity map [16] was compiled as the land necessary to sustain a
450 kg animal over an extended period of time, without degrading the soil or
vegetation. Grazing capacity is thus expressed as hectare per large stock unit
(ha LSU-1) and is relevant only for vegetation that is in relatively good condition.
The expectation is that the animal grows 0.5 kg per day from grazing alone that has a
digestible energy content of 55%. The grazing capacity map largely follows the
rainfall and evaporation map (Fig. 1) and varies greatly, from less than 2 ha LSU-1
in the east to 140 ha LSU-1 in the arid northwest. Stocking animals in excess of these
guideline values is therefore expected to degrade the vegetation and soil, whilst
stocking to these values on degraded land will further degrade the vegetation and
soil. Unfortunately communal farmers have little motivation, capacity, or infrastruc-
ture to manage their animal numbers to these norms [19]. This is one of the primary
drivers of vegetation and soil degradation in communal areas.
334 C. W. van Huyssteen and C. C. du Preez

Fig. 5 Grazing capacity map for South Africa, expressed as the number of hectares required to
maintain a large stock unit (ha LSU-1), based on data adapted from DAFF [16]

3 Land Use and Production

Climate, together with soil and vegetation, is the main determinant of the most
sustainable agricultural production practice (Fig. 6). Agriculture, in this context,
includes livestock and crop production. For crop production, climate and soil are the
main factors, whilst climate and vegetation are the primary factors for livestock
production.
About 101 × 106 ha land is used for agriculture, whilst an additional 1.4 × 106 ha
is used for commercial forestry (Table 1). Only 16.8 × 106 ha land in South Africa
can be considered suitable for cropping, of which 14.2 × 106 is deemed to be
commercial, whilst 2.6 × 106 ha is deemed to be communal. Of this, a total of
12.9 × 106 ha is under crop production, comprising 11.5 × 106 ha dryland and
1.4 × 106 ha irrigation [4].
Most of South Africa, about 84 × 106 ha, is only suitable for extensive grazing
(Table 1), primarily due to the high evapotranspiration and low rainfall. Commercial
livestock production, on 72 × 106 ha land, and communally livestock production, on
12 × 106 ha land, therefore comprise the greatest part of South Africa. Total livestock
numbers are estimated at 13.7 × 106 cattle, 21.1 × 106 sheep, and 1.9 × 106 goats
(Table 3) [4]. The numbers of poultry, ostriches, and pigs are smaller. A synopsis of
Agricultural Land Degradation in South Africa 335

Fig. 6 Main agricultural production activities per magisterial district in South Africa, based on data
adapted from the WWF [20]

the dominant production enterprise per magisterial district is given in Fig. 6. Land
tenure (commercial vs. communal), production enterprise, and climate impact dif-
ferently on agricultural land degradation and are also uniquely affected by the
consequences of this degradation. These aspects are addressed in the next two
sections.
The agricultural production areas discussed above resulted in the following
production, primarily for the 2018/19 production season: of the agronomic com-
modities (Table 2), maize occupied the largest area (2.60 × 106 ha), followed by soya
beans (0.71 × 106 ha), wheat (0.54 × 106 ha), and sunflower (0.52 × 106 ha).
Production followed a similar pattern, except for sugarcane, which has a much
higher moisture content at harvest. For livestock commodities (Table 3), the greatest
production was poultry, with 1,861,000 Mg, followed by beef (1,027,000 Mg), pigs
(260,000 Mg), and sheep and goats (170,700 Mg). In terms of gross value, cattle
production contributed the greatest fraction of the livestock commodities. In terms of
gross value, grapes and oranges constitute the predominant horticulture production
(Table 4), followed by apples, and then approximately equal contributions from
lemons and limes, pears, and grapefruit. Potatoes, green mealies (corn) and
sweetcorn, tomatoes, and onions contribute the greatest gross value for the vegetable
crops (Table 5).
336

Table 1 Summary of land use in South Africa and the individual provinces, as determined during 1991, based on data adapted from DAFF [21]
Farmland
Arable land
Potentially arable land utilisation Grazing land
Commercial Communal Commercial Communal Nature
Country and agriculture agriculturea Dryland Irrigation agriculture agriculturea Total conservation Forestry Otherb Total
provinces 106 ha
South Africa 14.19 2.55 11.54 1.36 71.99 11.93 100.67 11.79 1.43 8.43 122.32
Eastern Cape 0.64 0.53 0.41 0.19 10.17 3.47 14.82 0.62 0.13 1.49 17.06
Free State 4.19 0.04 3.86 0.14 7.39 0.15 11.79 0.27 0.00 0.91 12.94
Gauteng 0.44 0.00 0.38 0.03 0.39 0.00 0.83 0.23 0.02 0.79 1.89
KwaZulu- 0.84 0.36 0.70 0.13 2.60 2.73 6.53 1.38 0.47 0.78 9.15
Natal
Limpopo 1.17 0.53 0.50 0.16 5.98 2.86 10.55 1.16 0.07 0.02 11.96
Mpumalanga 1.60 0.14 1.61 0.13 2.89 0.35 4.98 2.33 0.55 0.32 8.18
Northern 2.41 0.00 0.03 0.19 29.09 0.00 29.54 4.29 0.00 0.45 36.33
Cape
Northwest 0.45 0.95 2.21 0.10 4.38 2.36 10.10 0.76 0.00 1.01 11.87
Western 2.45 – 1.84 0.29 9.11 – 11.56 0.73 0.19 0.45 12.94
Cape
a
In former homelands
b
For example, urban, mining, and industry
C. W. van Huyssteen and C. C. du Preez
Agricultural Land Degradation in South Africa 337

Table 2 Summary of the major agronomic commodities produced in South Africa, based on data
adapted from DALRRD [4]
Area planted Total production Gross value
Commodity Year 1,000 ha 1,000 Mg ZAR1,000 USD1,000a
Maize 2018/19 2,596 11,824 28,115,441 1,945,705
Wheat 2019 540 1,508 6,115,034 423,186
Grain sorghum 2018/19 50 146 437,693 30,290
Groundnuts 2018/19 20 22 170,856 11,824
Sunflower 2018/19 515 705 3,476,405 240,582
Soya beans 2019/20 705 1,243 5,836,634 403,919
Oats 2019 21 16 3,961 274
Barley 2019 132 345 3,040 210
Canola 2019 74 96 5,504 381
Sugar cane 2018/19 395 19,302 452 31
a
Average exchange rate for 2019: USD1 = ZAR14.45

Table 3 Summary of the major livestock commodities produced in South Africa, based on data
adapted from DALRRD [4]
Number
slaughtered Production Gross value
Commodity Year Number 1,000 (Mg) ZAR1,000 USD1,000a
Cattle 2017/ 12,800,000 3,311 1,027,000 47,144,495 3,262,595
18
Pigs 2017/ 1,454,000 3,134 260,000 6,407,960 443,457
18
Sheep 2019 19,389 6,145 170,700 12,356,802 855,142
Goats 2019 1,781
Poultry 2019/ N/A N/A 1,861,000 N/A N/A
20
a
Average exchange rate for 2019: USD1 = ZAR14.45

Table 4 Summary of the major horticulture commodities produced in South Africa, for the 2018/
19 production year, based on data adapted from DALRRD [4]
Total production Gross value
Commodity 1,000 Mg ZAR1,000 USD1,000a
Grapes 1,852,299 13,270,411 918,368
Oranges 1,774,397 10,891,097 753,709
Apples 856,642 5,794,644 401,013
Lemons and limes 473,197 3,741,627 258,936
Pears 414,000 2,987,160 206,724
Grapefruit 445,351 2,500,262 173,029
Peaches 151,841 1,167,657 80,807
Plums 58,815 918,408 63,558
Tangerines 53,230 228,703 15,827
Apricots 33,288 222,973 15,431
Prunes 1,786 1,756 122
a
Average exchange rate for 2019: USD1 = ZAR14.45
338 C. W. van Huyssteen and C. C. du Preez

Table 5 Summary of the major vegetable commodities produced in South Africa, for the 2019
production year, based on data adapted from DALRRD [4]
Production Average price Gross value
Commodity (1,000 Mg) (ZAR/Mg) ZAR1,000 USD1,000a
Potatoes 2,485 3,948 9,810,780 678,947
Green mealies and 394 22,123 8,716,462 603,215
sweetcorn
Tomatoes 558 7,020 3,917,160 271,084
Onions 724 3,854 2,790,296 193,100
Carrots 217 3,619 785,323 54,348
Pumpkins 265 2,204 584,060 40,419
Cabbage 161 2,481 399,441 27,643
Beetroot 83 3,761 312,163 21,603
Sweet potatoes 82 3,457 283,474 19,618
Green beans 22 11,694 257,268 17,804
Green peas 5 41,341 206,705 14,305
Cauliflower 11 12,371 136,081 9,417
a
Average exchange rate for 2019: USD1 = ZAR14.45

4 Soil Degradation

Production activities on range-, crop-, and forestland hugely impact soil quality and
vice versa. Soil quality can be defined as the ability of soil to function efficiently in
an ecosystem to sustain biologically production [22]. This guarantees that the
environment stays unblemished and the health of humans, animals, and plants is
advanced.
A decrease in soil quality results from a range of degradation processes that may
be categorised as biological, physical, and chemical (Fig. 7). These processes are
predominantly caused by anthropogenic activities and may influence one another.
For example, salinisation, especially by sodium salts, causes structural decay, which
often manifests in surface sealing, inducing reduced water infiltration, higher water
runoff, and thus the increased risk of water erosion. Similarly, acidification of soil
results in the concurrent release of aluminium and fixation of phosphorus – a double-
edged sword for plant growth and development, because phosphorus is essential,
whilst aluminium is toxic [24].
Generally, it is assumed that 25% of soil in South Africa is severely degraded by
some of the processes shown in Fig. 7. Information on the contribution by either
communal or commercial farming is, however, almost non-existent. It is generally
perceived that water erosion, followed by wind erosion, salinisation, and acidifica-
tion, is dominant depending on the soil type, climate, land tenure, and land use.
Agricultural Land Degradation in South Africa 339

Fig. 7 Summary of the biological, chemical, and physical processes of soil degradation, adapted
from Lal and Stewart [23]

4.1 Plinthite Formation

Impervious plinthite horizons occur in <1% of South African soils (Fig. 2), in the
more humid areas from central to north-eastern South Africa [8]. These layers
develop due to iron and manganese that are mobilised and accumulate during
wetting and drying cycles. The process is natural and reduces the soil’s depth for
root growth and development and cannot be prevented. A shallower effective rooting
depth inhibits the capacity of soil to provide adequate water and nutrients for
optimum plant growth and root development. Plinthite formation is not considered
to be of concern in South Africa because most soils would dry out naturally in the
recent as well as geological past, leading to the natural induration of those horizons
that are prone to do so [25].

4.2 Surface Crusting

The current knowledge on South African topsoil crusting was recently reviewed by
Laker and Nortje [26]. They established that soil crusting is a widespread and severe
problem. Soil crusting leads to reduced water infiltration, increased water erosion,
and reduced seedling emergence. Unlike subsoil compaction, anthropogenic soil
crusting is not restricted to croplands but is also problematic in the rangelands used
for livestock and game ranching. Little is known about topsoil crusting on forestland,
but it is expected to occur on some soils, especially when the trees are still young and
a litter layer is either absent or miniscule.
Crusts are divided into two groups, namely mineral crusts and biological crusts
[27]. Mineral crusts develop from the disintegration of soil structure, resulting in a
340 C. W. van Huyssteen and C. C. du Preez

dense soil surface layer. In this process, the clay at the surface disperses during
wetting and ultimately blocks soil pores upon drying [28]. South African soils most
vulnerable to this crusting have clay fraction dominated by smectite, exchangeable
sodium >2%, organic matter content <0.2%, or Ca:Mg ratio <1 [26, 29]. However,
soils not having these properties can also form crusts mainly due to raindrop impact
[28], especially when the protective vegetation layer is removed. Mineral crusts
hamper seedling emergence and water infiltration. The extent of mineral crusting in
agriculture and forestry is unknown. This can probably be attributed to the fact that
mineral crusts are mostly detected in local patches of the landscape. Practices to
combat the detrimental impact of mineral crusting are maximum residue retention
under crop farming and retention of good vegetative cover in livestock and game
ranching [30].
Less is known about the biological crusts that form by a close relationship
between soil organisms and soil particles [26]. Soil organisms thus aiding in crusting
comprise inter alia of bryophytes, lichens, bacteria, micro-fungi, green algae, and
cyanobacteria that live in or on top of the soil [31]. Biological soil crusts can in semi-
arid and arid areas protect sandy soil against wind erosion. Serious biological
crusting is often observed in irrigated orchards under all climates, mainly in the
shaded areas under tree canopies [26]. The occurrence of this crusting can therefore
also be expected in forestland. Although not reported, biological crusting is also
probable in the cooler, wetter rangelands. Biological crusts protect soil against
erosion, but physical disruption counters this benefit [26]. Animal trampling in
rangelands, tillage of croplands, and either tree planting or harvesting in forestlands
may damage these crusts, thus increasing the risk of soil to either wind or water
erosion.

4.3 Subsoil Compaction

Many South African soils, especially the sandy variants of the red, yellow, and grey
plinthic and apedal soils (Fig. 2), have dense subsoils in their natural state [32]. In
these hard-setting soils, bulk densities of 1,650 kg m-3 are not uncommon. How-
ever, subsoil compaction is induced by mechanisation in about 2 × 106 ha cropped
soils [24]. Maize and wheat are mainly grown on these soils that usually have well-
rounded and well-sorted sand, with low clay content [33]. The following depth zones
are observed in these soils [34]:
Zone 1 (<20 mm): A dense crust, but that is not always present.
Zone 2 (20–150 mm): Relatively loose cultivated layer within the upper part of the
plough layer.
Zone 3 (150–250 mm): Compacted layer in the lower part of the plough layer.
Zone 4 (250–450 mm): Compacted layer within the sub-plough layer.
Zone 5 (>450 mm): Naturally compacted layer in soils prone to compaction.
Agricultural Land Degradation in South Africa 341

In their review, Laker and Nortje [32] expanded on the soil factors (clay miner-
alogy, particle size distribution, and organic matter content) as well as management
factors (uncontrolled vehicular traffic, high tyre pressure, cultivating and traversing
on wet soil, and secondary tillage operations) contributing to subsoil compaction.
They also address the effects and consequences of subsoil compaction (weak root
and top growth and development, reduced uptake of nutrients, poor water use
efficiency, reduced hydraulic conduction, and aggravation of root diseases). Deep-
tine tillage and controlled wheel traffic are advocated to relieve subsoil compaction
[34]. It is further recommended to prevent cultivation on wet soils and to use low tyre
pressures to avoid subsoil compaction [32]. Conservation agriculture that comprises
inter alia of minimum soil disturbance (no-tillage) and maximum residue retention
(mulching) can be considered to counter subsoil compaction, but reliable research
data are currently lacking – an aspect that requires thorough investigation. Subsoil
compaction manifests in 15–30% lower crop yields due to ineffective water and
nutrient use resulting from poor root growth and development.
Compaction of subsoils is not confined to croplands only. Animal trampling and
off-road driving also induce subsoil compaction in rangelands for either livestock
farming [35] or game ranching [36] farming. Forestry soils in South Africa are also
very susceptible for subsoil compaction by heavy equipment used during
harvesting [37].

4.4 Structural Decay

Reliable information on structure decay is lacking in South Africa, except where it is


related to organic carbon gain or loss [38]. In ploughed soils, organic matter
decomposition is accelerated through better aeration and less physical protection,
leading to aggregate disintegration [39]. Structural decay results in surface crusting
that reduces water infiltration and negatively impacts on seedling emergence. This
structural decay coincides with loss of organic matter and reaches equilibrium after
17 years for the 2.8–8 mm fraction. The 2–2.8 mm aggregate fraction showed a
similar tendency, with a slightly delayed equilibrium. Plant debris is stabilised as
lignin in these two aggregate fractions [40]. Conversion of ploughed soil to perennial
secondary pasture sequestered organic carbon by rebuilding aggregates greater than
2 mm. Organic carbon reached 60% after 25 years in these large grassland aggre-
gates [41]. Regeneration of soil aggregates by this management practice is probably
not economically viable for most farmers, because net income from rangeland is
considerably lower than from cropland. However, farmers can consider conservation
agriculture through minimal soil disruption and the maximal retention of organic
residue as an alternative management practice, like no-tillage. Currently, less than
5% of cropped soils in South Africa is under conservation agriculture. Convention-
ally ploughed agriculture is practised on the remaining cropped soils. Structural
decay to varying degrees may thus potentially occur in 12.3 × 106 ha ploughed land.
342 C. W. van Huyssteen and C. C. du Preez

4.5 Erosion

This degradation process was reviewed by Scotney [42] and almost 25 years later by
Laker [43]. Both reviews cover erosion extensively and raised the concern that soil
loss through erosion exceeded soil formation. Du Plessis [44] estimates that about
0.3 Mg soil is formed per hectare per year, whereas soil removal from cropland and
rangeland is predicted at 1.7 and 3.0 Mg ha-1 year-1, respectively [45]. In eastern
South Africa, where the rainfall is higher, water erosion predominates. Conversely,
in the arid western parts wind erosion predominates. However, no clear boundary
line exists between these parts, because several factors (e.g. climate, topography, soil
type, and land use) influence the intensity of these two erosion processes. Erosion
contributes to air and water pollution, resulting in deteriorating environmental
quality.

4.5.1 Wind Erosion

About 25% of South African soil is vulnerable to wind erosion [46], whilst Du
Plessis [44] estimates that 2.2 × 106 ha land is severely affected by this degradation
process. The sandy soils of the cropped rain-fed fields in western South Africa are
very susceptible to wind erosion. This comprises the largest part of the Northwest
Province, the western Free State, and the western and northern parts of the Northern
Cape Province. The sandy coastal soils are also very susceptible to wind erosion.
Conventional tillage in the western Free State, with a fallow period of 10 months,
resulted in 59 Mg ha-1 year-1 soil loss through wind erosion [47]. However, under
mulch tillage soil loss was only 11 Mg ha-1 year-1. The reduction of the fallow
period from 10 to 5 months, when conventional mouldboard ploughing is practised,
decreased wind erosion from 59 to 23 Mg ha-1 year-1.

4.5.2 Water Erosion

Water erosion varies greatly, depending on the prevailing conditions. On the same
soil type in the central part of the Free State Province, soil loss increased from 0.4 to
2.6 Mg ha-1 year-1 when the condition of rangeland deteriorated from good to poor
[48]. Water erosion [18] is caused mainly by overstocking, which manifests in poor
rangeland conditions. Other noteworthy factors include soil compaction in trampled
areas, reduction of water infiltration, and increased runoff.
The production of maize, cassava, and pineapple results in uncovered soil under
conventional tillage and is hence often also prone to great soil losses [49]. For maize,
the situation is worsened by the crop being planted extensively under conventional
tillage, with measured soil losses around 25 Mg ha-1 year-1, which can be more
than 60 Mg ha-1 year-1 in severe cases. However, soil loss through water erosion
under conservational tillage is measured at less than 2 Mg ha-1 year-1 when >30%
Agricultural Land Degradation in South Africa 343

residue cover is sustained. These amounts are affected by inherent factors like
rainfall intensity, soil stability, topography [43].
In the Eastern Cape Province, pineapple farming is notorious for loss of soil
through water erosion. Losses as high as 123 Mg ha-1 year-1 were measured.
Chopped old pineapple plants left as mulch and placing whole pineapple plants
inverted in the middle of the ridges reduced soil loss noticeably [50], from over
48 Mg ha-1 to less than 2 Mg ha-1 on a 10% slope [51, 52].

4.6 Fertility Decline

The period from 1953 to 1978 was highlighted by Grant [53] in her review for the
noteworthy upsurge in cultivation. She attributed this to an increase in the intensive
usage of stronger tractors and larger ploughs and to improved yields from better
hybrid seed. Little had been done concerning the maintenance of soil fertility. This
trend still continues, as concluded by Barnard and Du Preez [54].
From 1960 to 1980, 60 Gg nitrogen and 470 Gg potassium were removed
annually from South African cropland, more than was applied [55]. At the same
time, 90 Gg more phosphorus was applied annually than removed. This trend
possibly still continues, since sales of nitrogen, phosphorus, and potassium peaked
in 1980, at 520, 240, and 130 Gg, respectively. Sales then decreased and for the past
three decades averaged 400 Gg nitrogen, 80 Gg phosphorus, and 90 Gg potassium
annually [56, 57].
Since 1980, the quantity of nitrogen applied relative to the phosphorus and
potassium applied increased [55–57]. At the same time, farmers depended heavily
on the soil nitrogen and potassium reserves, whilst reserves of soil phosphorus were
built up. In their natural state, South African soils contain low phosphorus and high
potassium reserves. Thus, nitrogen was mainly released through organic matter
oxidation [54]. These deductions are confirmed by limited research on soil analytical
laboratory data [58].
Comparable information for calcium, magnesium, and sulphur is lacking. Indi-
cations are, however, that these nutrients are removed to a greater extent from
cropland than being applied. Zinc is the only micronutrient regularly applied to
South African croplands [54, 58, 59], because maize is the major crop being
cultivated. Maize is sensitive to insufficient zinc supply from the soil, if not
supplemented.
Barnard and Du Preez [54] reviewed research on essential plant nutrients and
other elements such as sodium, silicon, and selenium and that were deemed to be
adequate. However, the ascribed soil fertility benefits of different crop production
management practices led to renewed interest in conservation agriculture under
commercial [60–63] and communal [64, 65] farming.
For commercial farmers, excellent fertilisation guidelines exist for numerous
crops [66], supplemented by several site-specific studies [67–70]. This is not applied
for communal farmers in the Limpopo Province, and Ramaru and others [71]
344 C. W. van Huyssteen and C. C. du Preez

therefore developed a soil fertility management framework for these farmers. The
framework was successful in educating a community on the importance of
maintaining and improving soil fertility for sustainable cropping. All other eight
provinces can also benefit from this innovation.
The soil fertility status of “homefields” (fields close to the homestead) and “out-
fields” (fields further from the homestead) in two KwaZulu-Natal communal farming
communities was investigated by Roberts et al. [72]. In Obanjaneni, the median-
adjusted phosphorus was 4 Mg L-1 for “outfields” and 14 Mg L-1 for “homefields”.
The median-adjusted phosphorus was 11 Mg L-1 for “homefields” and 3 Mg L-1 for
“outfields” in the Valley of Thousand Hills. Comparable tendencies were observed
for potassium, implicating that these communal farmers applied two different
fertilising systems in their fields. Thus, individual fields must be sampled for
fertiliser recommendations.
Mkhabela and Materechera [73] observed that communal farmers in the
KwaZulu-Natal midlands use mainly manure to increase the fertility status of
soils. The available manure is, however, insufficient for this purpose. The farmers
prefer to use cattle manure instead of chicken manure, notwithstanding the greater
benefit of the latter. Traditionally, these communal farmers believe cattle manure to
be the best. Complementary organic and inorganic fertiliser use ought thus to be
encouraged to improve soil fertility in the long term.
Two-thirds of communal farmers in the Northwest Province apply manure to their
crops [74]. Manure from cattle is favoured, followed by sheep, goat, and chicken
manure. The use of manure was positively influenced by the size of farm, accessi-
bility of extension services, education, and effort involved. None of manure avail-
ability, land ownership, and herd size affected the choice not to use manure. Manure
should thus be used complementarily to chemical fertiliser.
Soil fertility also plays an important role in South Africa’s ecology, particularly
with regard to the dry savanna [75]. The importance of soil fertility for fynbos
vegetation was estimated by Richards et al. [76]. The effects of rangeland condition
on soil fertility properties of grassland [77] and savanna [78] biomes were also
examined. Swanepoel et al. [79, 80] investigated the fertility status of kikuyu-
ryegrass pasture under minimum tillage in the Western Cape Province.
In the Eastern Cape Province, the transformation of succulent thicket induced a
decrease of soil fertility [81]. A distinct soil fertility spatial pattern was observed for
intact thicket, with nitrogen and phosphorus accumulated below the perennial
shrubs. This caused a homogenisation of nitrogen and phosphorus. As a conse-
quence, soil fertility generally decreased in the landscape.
Acacia tortilis and Burkea africana plant populations occur as patches on sandy
soils with similar parent material at Nylsvlei in South Africa. Concentrations of some
macronutrients (phosphorus, potassium, calcium, and magnesium) under the Acacia
communities were 10–100 times greater than under the Burkea communities. Almost
similar sodium levels were observed in both plant communities. Blackmore et al.
[82] attributed the differential accumulation of these macronutrients to the historic
Tswana tribe homesteads, rather than to landscape geomorphology.
Agricultural Land Degradation in South Africa 345

In the Eastern Cape Province, largely consisting of tribal land, important micro-
nutrient deficiencies that might cause oesophageal cancer were observed [83]. In the
KwaZulu-Natal Province, osteoarthritis was noted by Ceruti et al. [84]. A lack of
iron was recognised as impacting children’s intellectual development, impacting
human productivity negatively. Iron deficiency may also be the cause for scholarly
losses, increased mortality, and higher maternal deaths [85].
Lower trace element concentrations were reported in soils of resource-deprived
farmers of Mpumalanga Province, compared to soils of the adjacent rangelands
[86]. A comprehensive study of South African soils showed that many soils have a
natural deficiency in trace elements [87]. On the other hand, toxic high trace element
concentrations were observed in some parts of South Africa that may manifest in
health problems. These areas typically have low inherent fertility and typically
receive minimal fertiliser input under subsistence and communal land tenure
systems.

4.7 Elemental Imbalance

Elemental imbalance refers to the excessive accumulation or depletion of elements


(e.g. acidic cations, salts, or pollutants) to such an extent that it negatively impacts
soil properties, vegetative growth, or human health.

4.7.1 Acidification

Prolonged tillage and excessive ammonium fertiliser addition contribute to the


acidification of cultivated soil. This is probably the most important reason for
decreased soil fertility from anthropogenically induced soil acidification resulting
in plant nutrition problems. For instance, lower nitrogen and sulphur concentrations,
caused by lower organic matter, caused a decrease in plant accessibility of phos-
phorus and molybdenum, an increase in metal cations’ solubility, often resulting in
copper and zinc leaching, and occasionally iron toxicities [54].
The moderately acidified natural land, with pH(KCl) between 4.5 and 5.5, is
estimated at 16 × 106 ha. Another 5 × 106 ha is strongly acidified and has pH(KCl)
<4.5. Most of these soils occur in the higher rainfall southern and eastern parts of
South Africa [88].
The quantification of soil acidification by anthropogenic activities is a larger
challenge. It has, however, been estimated in the 1990s that of the 12.9 × 106 ha
cultivated land in South Africa, approximately 4 × 106 ha topsoil and 2 × 106 ha
subsoil are already severely acidified. The consequences are lower yields and less
income for farmers [88]. Management practices aim to rectify anthropogenic soil
acidity by liming to increase pH to satisfactory levels or to decrease exchangeable
acidity to tolerable levels. The most appropriate approach depends largely on
whether the amount of lime required to attain these levels is affordable to the farmer.
346 C. W. van Huyssteen and C. C. du Preez

At a symposium entitled “Soil Acidity Initiative”, Beukes [89] reviewed


South Africa’s soil acidity dilemma, Farina [90] covered subsoil acidity, and
Haumann [91] examined soil acidity management. Two relevant papers were also
presented by Fey [92] and Haynes [93] at the fifth International Plant-Soil Interac-
tions at Lower pH Conference. They addressed the negative effects of high soil
acidity in forestry, agriculture, and natural resources, emphasising the obstacles
which hinder the progression in nutrient use efficacy as outlined by Van der
Merwe et al. [94]. From these contributions, it can be concluded that acidification
of soils in areas with more than 500 mm annual rainfall is a matter of concern,
especially in croplands. The growth and development of most crops are thus
negatively affected if this soil degradation process is not effectively managed
through liming.

4.7.2 Salinisation

Irrigation degrades the soil through waterlogging, salinisation, and sometimes


sodification. This is ascribed to low soil suitability, a decline in water quality, poor
management, and economic pressure [95].
Soil salinisation is a problem on the 1.4 × 106 ha land irrigated in South Africa. Of
this, an estimated 140 × 103 ha irrigated land is affected by salinisation [44]. Approx-
imately, 40 × 103 ha is permanently salinised, whilst about 80 × 103 ha is temporarily
affected by salinisation [96]. Reliable information on salinisation in rangelands is
lacking. However, small patches of soil salinisation are found in the drier western
rangelands. These areas are established naturally, and the vegetation is normally
adapted to growth in these conditions.
Nell and Van Huyssteen [97] used soil profile and other data to map primary
salinity and sodicity of land. Based on this study, 0.4% is strongly saline, 5.1%
saline, 1.4% moderately saline, and 23% is slightly saline, whereas 0.4% is sodic and
6.3% alkaline saline-sodic. The area of land affected by primary salinity and sodicity
is therefore not as abundant as observed in other dry areas.
It is principally soils poorly suited to irrigation that are affected by salinity. These
soils’ salinity is remarkably constant with depth and rapidly recovers upon artificial
drainage. A larger danger [67, 98] is, however, the constant reuse of poor-quality
irrigation water. The challenge should therefore rather be to prevent the prior
salinisation of irrigation water.
Due to marked water quality deterioration downstream, water logging and con-
sequential sodicity and salinity of irrigated soils alongside the lower Vaal River and
its tributaries were investigated [99–102]. Electrical conductivity of the saturated
extracts from well-drained sandy soils under centre pivot irrigation has low salinity
(50–100 mS m-1); sandy soils with water tables developed salinity, whilst salt and
sodium accumulated in clay soils due to their incapacity to leach. Thus, the warning
of Du Plessis [98] that limited drainage could cause root zone salinity and sodicity of
irrigated soils is endorsed. Currently, it is acknowledged that crops can tolerate
greater root zone salinity levels than previously considered possible.
Agricultural Land Degradation in South Africa 347

In the Breede River, the declining water quality is also of concern since it might
lead to the salinisation of irrigated viticulture soils in the Western Cape Province
[101]. Full and supplementary irrigation of vineyards with 75–100 mS m-1 electrical
conductivity, combined with different frequencies and kinds of irrigation, was
investigated by De Clercq et al. [103]. Cultivars differed in tolerance to soil
salinisation, which was related to the soil depth, season, and irrigation kind. For
these situations, saturated soil extract electrical conductivity varied from 50 to
225 mS m-1. Subsurface drip, however, caused larger build-up of salt at the surface
than surface water application.

4.7.3 Pollution

Non-point source pollution results from hazardous wastes from cropping such as
fertilisers and pesticides [95, 104]. One of the most polluted rivers in South Africa is
the Elephants River in the Mpumalanga Province. This is because 24% of the
catchment above the Witbank Dam is monocropped with maize [105], threatening
particularly fish eagles [106]. From the estimated 118 Mg nitrogen released from the
Witbank Dam catchment, less than 20% stems from agricultural activities. A sub-
stantial higher percentage of discharged phosphorus, viz. 53%, originated from
agriculture. The pollution of irrigated water by nitrogen and phosphorus does not
usually pose a serious threat for cropping, although excessive nitrogen might result
in vigorous vegetative growth and hence delayed ripening.
Amending soil with effluent and organic waste from piggeries and feedlots is
usually advantageous, since it improves soil quality. Excessive application may,
however, increase copper, zinc, and mercury concentrations in the water and soil. No
significant concentration increases of other metals were measured.
Surplus water from industrial effluent is often irrigated on pastures to dispose
thereof through evapotranspiration. For example, the SASOL coal gasification
factory effluent contains fluorine and boron levels that could be detrimental for
plant growth and development [95].
Coal-burning power generation plants produce about 26 × 106 Mg pulverised fly
ash, which is frequently amended to soil. Toxic elements such as arsenic, barium,
bismuth, manganese, copper, and vanadium may thus end in soil through the fly ash.
The high pH of the ash counters the possible pollution hazard of these elements by
reducing their solubility and mobility [95]. Raised levels of cadmium, antimony, and
mercury were found close to a coal-fired station in Bloemfontein [107]. Some local
sites were polluted by arsenic, but none by selenium. The elevated levels of the
elements were attributed to ash of the power station as well as to other industries. In
most cases, soil pollution is related to the source and thus occurs in small regions,
hence facilitating the monitoring and management thereof.
348 C. W. van Huyssteen and C. C. du Preez

4.8 Organic Matter Change

No systematic study exists on the spatial distribution of soil organic carbon in


South Africa, probably due to either poor insight or financial constraints. However,
Barnard [108] used modal profile data from the land type survey [8] to produce a
generalised map of soil organic carbon content for the upper 300 mm of virgin soils
and concluded that only 4% of soils contain more than 4% soil organic carbon,
whilst 58% contain less than 0.5% soil organic carbon. Rantoa et al. [109] used these
modal profile data to statistically analyse soil organic carbon contents in master
horizons. Reported results are in accord with what can be expected from the master
horizons’ definitions [110].
The soil organic carbon content and quality change due to field crop production in
South Africa were recently reviewed [38]. Chronosequence studies show that
converting rangeland to conventionally tilled cropland leads to >50% decline in
soil organic carbon after 50 years, whilst establishing perennial pasture leads to only
a 60% recovery in soil organic carbon after 30 years. Conservation agriculture,
which includes maximum residue retention, minimum soil tillage, and proper crop
rotation, maintains and in some instances increases soil organic carbon levels. Long-
term trials [111] show that an increase in soil organic carbon is site-specific and
justifies more research.
Traditionally grassland or forest soil is converted for the production of sugarcane.
On average, the crop is replanted after eight ratoons. Before harvesting, the crop is
burnt to remove leafy biomass that does not contain sucrose. The remains after
harvesting are mostly left on the surface of the soil. This depleted soil organic carbon
by 25–70% [112, 113]. Management options suggested to protect soil organic
carbon are planting of green manure crops for more biomass production, harvesting
of green cane with residue retention, and practising minimum tillage in controlled
lanes [114].
Swanepoel et al. [115] report that natural fynbos vegetation in the southern Cape
coast is often cleared to establish kikuyu pasture for milk production. This change
includes conventional tillage, fertilisers, and irrigation which all enhanced a decline
in soil organic carbon. Research by Swanepoel et al. [79, 80, 116] shows that this
decline can be countered by reduced tillage in combination with kikuyu-ryegrass
mixed pasture.
The change of soil organic carbon due to afforestation of grassland in the
Weatherley catchment of the Eastern Cape Province was investigated. Eight years
of forestry consisting of Pinus elliottii and Eucalyptus nitens reduced soil organic
carbon stock significantly from 47.6 to 38.8 Mg ha-1, whereas P. patula marginally
improved soil organic carbon stock from 42.8 to 48.6 Mg ha-1 [117].
Rangeland scientists estimate that 66% of the country’s 84 × 106 ha rangeland is
moderately to severely degraded. This results in a decline of soil organic carbon
content as observed by Du Preez and Snyman [118]: 25% in the surface 200 mm of a
sandy loam soil, 15 years after vegetation in good condition, was transformed to
vegetation in poor condition. In this period, 4,200 kg carbon ha-1 was lost in excess
Agricultural Land Degradation in South Africa 349

from the poor vegetation’s soil than from the good vegetation’s soil. Subsequent
studies [29, 77, 78] confirmed these results [118].
Game ranching, like livestock farming, also causes a decline in soil organic
carbon when the rangeland condition deteriorates. Efficient management of range-
land for both kinds of farming is therefore essential for sustainable production [17].

4.9 Organism Change

Despite limited information, Haynes and Graham [119] reviewed the impact of land
use change and soil disturbance on organism diversity and activity in South African
soils. Under sugarcane, a decrease in soil organic carbon resulted in a decrease of
microbial biomass carbon, microbial quotient, fluorescein diacetate (FPA) hydro-
lytic activity, and basal respiration [112]. By using grassland as a reference, kikuyu
pasture had increased soil organic carbon and therefore soil basal respiration as well
as microbial biomass carbon, whilst maize and sugarcane had clear decreases
[120]. In a long-term trial where pre-harvest burning against trash preservation
was investigated, Graham et al. [121] found that 60 years of green cane harvest
and/or NPK fertilisation improved the yield that manifested in greater soil organic
carbon and hence microbial biomass carbon, basal respiration, and microbial quo-
tient for the 0–300 mm depth.
For the University of Pretoria experimental farm, Belay et al. [122, 123] report
that 60 years of annual fertiliser and manure applications manifested in higher soil
organic carbon, microbial biomass carbon, and bacterial counts than in the control
plots. The nitrogen, phosphorus, and potassium fertilised plots had higher microbial
biomass carbon, and counts of bacteria, fungi, and actinomycetes than the manured
plots.
Under annually tilled ryegrass pastures in the Eastern Cape Province, Milne and
Haynes [124] reported lower soil organic carbon, extractable carbon, microbial
biomass carbon, fluorescein diacetate hydrolysis, organic ammonification, and
basal respiration rates than under permanent kikuyu pastures. Similar results are
observed in podzols with minimum-till, irrigated kikuyu-ryegrass pastures after
19 years [115].
Fynn et al. [125] investigated the scheduling and frequency of grassland burning
on soil organic carbon as well as microbial activity. Generally, burning schedule had
little effect, but burning frequency led to decreases of soil organic carbon in the
0–20 mm soil depth and microbial biomass carbon increases in the 20–40 mm soil
depth. More studies of this nature are suggested to obtain better insight of the soil
organic carbon and microbial biomass carbon responses to grassland burning.
In rangeland environments, the drier sandy savanna had lower enzyme activities
and phospholipid fatty acid contents than the wetter clayey grassland under both
commercial and communal management practices [126]. This is probably due to the
soil texture that plays a vital part in preserving microorganism communities. On the
other hand, particular enzymes were elevated in the sandy savanna ecosystem, after
350 C. W. van Huyssteen and C. C. du Preez

the microbial activity was standardised against soil organic carbon and microbial
biomass carbon, indicating more efficient microbe functioning in the sandy
ecosystem.
In a long-term trial where pre-harvest burning with trash management was tested
on sugarcane soils, Graham et al. [121] found clear contrasts in phospholipid fatty
acid profiles. This suggested community structural differences between sugarcane
soils under pre-harvest burning with trash management and less differences due to
long-term fertilisation. Metabolic variety displayed a comparable trend. A higher
soil organic carbon content led not only to an escalation in aerobic microorganism
numbers but also to a proliferation of the Gram-negative bacteria, whereas, under
unfertilised trash preservation, the fungi to bacteria ratio increased.
Earthworms in agricultural soils of the KwaZulu-Natal Province are mostly alien
species [120, 127]. Species from South America, India, and West Africa occur in the
northern, tropical parts, whilst European species cohabit with Indian and West
African species in the colder midlands. The earthworm quantities and biomass are
related to soil organic carbon, extractable carbon, and microbial biomass carbon.
Earthworm species and quantities are higher in permanent kikuyu pastures and lower
in cropped fields. The earthworm numbers were improved by the change of maize
fields from conventional mouldboard ploughing to no tillage and by the change of
sugarcane fields from pre- harvest burning to green cane harvesting. Low earthworm
quantities were observed in forest soils, probably because of the trash’s low
palatability.

4.10 Pathogen Increase

Soil-borne diseases often result from improper land and soil management, particu-
larly with monocropping. Prudent crop rotation systems can thus restrict soil-borne
diseases like “take all” (Gaeumannomyces graminis) in wheat, “root rot” (Fusarium
spp.) in maize, and “common scab” (Streptomyces spp.) in potato production.
Common scab is usually more severe in the neutral to alkaline soils and can therefore
be controlled by managing soil pH [128]. Crop rotation is also effective in restricting
nematode numbers in maize cultivated soils. Pathogen increase results in lower yield
of lesser quality, increases the production costs, thus resulting in lower profitability,
and is therefore mainly of concern in the intensively cropped areas (Fig. 6).

5 Land Degradation

This appraisal of land degradation in South Africa reflects the current state of soil
and vegetation degradation, individually and combined, as presented previously.
Hoffman et al. [129] conducted 34 workshops across South Africa, which captured
the opinions of 453 agricultural extension officers as well as resource conservation
Agricultural Land Degradation in South Africa 351

Fig. 8 Soil degradation index for South Africa, modelled per magisterial district based on data
adapted from Hoffman et al. [129]

technicians on inter alia the extent of soil and vegetation degradation in each of the
367 magisterial districts. Data from these workshops were augmented by case
studies, selected to represent “worst-case” and “best-case” scenarios. The question-
naire used largely followed the procedure set out in the WOCAT manual.
The collected data were used to compile a soil degradation index (Fig. 8) and a
vegetation degradation index (Fig. 9). These two indices were then joined into a
combined soil and vegetation degradation index (Fig. 10). The soil degradation
index incorporated inter alia soil loss due to erosion, soil covering (due to
overblowing), nutrient mining, acidification, and pollution. The vegetation degrada-
tion index incorporated loss of vegetative cover, transformation of vegetation com-
position, alien plant invasion, bush encroachment, deforestation, and “other”. The
combined soil and vegetation degradation index was calculated as the sum of the soil
degradation and vegetation degradation indices [129].
Hoffman et al. [129] concluded that soil degradation (Fig. 8) is of greater concern
in the KwaZulu-Natal, Northern, and Eastern Cape Provinces and of lesser concern
in the Free State, Western Cape, and Northern Cape Provinces. In general, soils in
communal areas were far more eroded than soils in commercial areas. Sheet, rill,
gully, and donga erosion dominated, followed by wind erosion, whilst salinisation
and acidification occurred only in selected areas. Some forms of soil degradation are
352 C. W. van Huyssteen and C. C. du Preez

Fig. 9 Vegetation degradation index for South Africa, modelled per magisterial district based on
data adapted from Hoffman et al. [129]

decreasing, especially in commercial areas. These decreases were attributed to farm


planning, conservation works, legislation, education, planning, reduced stock num-
bers, and conversion to game ranching. Increased soil degradation, especially in
communal areas, was attributed to increased population, lacking infrastructure and
planning, poor education, poor runoff control, increased stock numbers, and the
diversification of grazing animals.
Regarding vegetation degradation (Fig. 9), the Northern and KwaZulu-Natal
Provinces arose as the provinces of greatest concern. The Northern Cape Province
also ranks relatively high, largely as a result of rangeland degradation due to the
spread of Prosopis species. Grazing lands in the communal areas were regarded to be
about twofold more degraded when compared to the commercial areas. Rangeland
degradation of concern included decreasing vegetative cover, alteration of species
composition, alien plant invasion, bush encroachment, deforestation, and “other”.
Decreasing vegetative cover and alteration of species composition were highlighted
as of greatest concern in more than half of the magisterial districts of South Africa.
Rangeland degradation through alien plant invasion was also widely recognised,
whilst bush encroachment and deforestation were identified in only a few magisterial
districts [129].
Agricultural Land Degradation in South Africa 353

Fig. 10 Combined soil and vegetation degradation index for South Africa, modelled per magiste-
rial district based on data adapted from Hoffman et al. [129]

The combined soil and vegetation indices (Fig. 10) highlight the communal areas
(former Transkei, Ciskei, and KwaZulu-Natal) of the eastern escarpment as amongst
the most degraded in South Africa. Communal areas in the Northern and North West
Provinces, together with several commercial areas in the Northern Cape and Western
Cape Provinces, were also shown to be severely degraded, but due to differing
causes. Communal areas are often significantly more degraded than commercial
areas, even though these areas may occur in the same ecosystem, adjacent to each
other. These differences were therefore mainly attributed to the different land tenure
systems. Finally, the Northern and KwaZulu-Natal Provinces emerged as the most
degraded provinces, followed by the Eastern Cape and North West Provinces, whilst
the Gauteng and Free State Provinces were the least degraded [129].

6 Summary

South Africa is an arid country that receives about half of the global average
precipitation per year. This precipitation occurs mainly in the form of high-intensity
rainstorms that increases the risk of water erosion on the steep slopes created by
354 C. W. van Huyssteen and C. C. du Preez

continental uplift. The extensive occurrence of clayey, duplex soils further increases
the risk of water erosion. South Africa has unique vegetation, grouped into nine
biomes. One of these, the Fynbos Biome is endemic to South Africa. Each of these
biomes has unique vegetative production, palatability, and animal nutrition and
therefore has unique management challenges and degradation potentials.
Only about 14% of South Africa’s land surface is arable, with the rest only suited
to forestry (1%) and extensive grazing (83%). The risk for land degradation is thus
determined by this differentiation in land use. Another overruling factor is the
different land tenure systems, with 71% being under commercial and 12% under
communal agricultural production.
Plinthite formation is not of great concern in the South African agricultural
environment, whilst surface crusting is. Mineral crusts occur extensively, especially
on smectite-rich clayey soils, with >2% sodium and <0.2% organic matter, where
the vegetative cover has been removed through ploughing or overgrazing. Subsoil
compaction of cropped sandy soils that inhibits root growth is of concern on about
2 × 106 ha, whilst soil structure decay, due to organic matter decline, potentially
occurs in 12.3 × 106 ha ploughed land.
Wind erosion is of concern on about a quarter of South African soils, with about
2% severely affected. Soil loss through water erosion varies from <1 to 60 Mg ha-
1
year-1 in severe cases and is largely resultant from inappropriate land use practices
(e.g. overgrazing and conventional tillage) on susceptible soils. Fertility decline of
cropland is worrying since 60 Gg nitrogen and 470 Gg potassium were removed
annually, in excess of what was being applied. This decline is much greater for
communal than for commercial farmers. About 16 × 106 ha of land is considered to
be natural acidic, whilst an estimated 12.9 × 106 ha of cultivated land has been
anthropogenically acidified.
Salinisation occurs on 1.4 × 106 ha land irrigated in South Africa, whilst sodicity
is not of great concern. Water pollution though nitrogen and phosphorus occurs
extensively in selected rivers in South Africa; however, agriculture was shown to be
responsible for only 20% of the nitrogen and 53% of the phosphorus, with the rest
emanating from human effluent and other industrial sources. South African soils are
naturally low in organic matter, with only 4% of soils containing more than 4% soil
organic carbon, whilst 58% contain less than 0.5% soil organic carbon. Conven-
tional tillage further decreases this by more than 50% after 50 years of cultivation.
Based on these degradation processes and as estimated by stakeholders, soil
degradation in the KwaZulu-Natal, Northern, and Eastern Cape Provinces is of
greatest concern, whilst it is of lesser concern in the Free State, Western Cape, and
Northern Cape Provinces. Water erosion, followed by wind erosion, salinisation, and
acidification, was identified as the major degradation processes and more so in the
communal than in the commercial areas. Rangeland degradation was of greatest
concern in the Northern, KwaZulu-Natal, and Northern Cape Provinces. Once again,
the communal areas were about doubly more degraded than the commercial areas.
In conclusion, land tenure as well as inappropriate land use and management
practices within the unique ecosystems emerged as primary drivers for agricultural
land degradation. On a nation scale, this degradation does, however, seem to be
Agricultural Land Degradation in South Africa 355

restricted to certain geographical areas, expediting the educational, planning, and


legislative efforts to address these.
Future research should consider the impact of conservation cropping on land
degradation, the improvement of conservation cropping systems, the impact of high
density versus conventional grazing systems on grazing land degradation, and lastly
soil health, soil organisms, and soil pathogen interactions.

References

1. Du Preez CC, Kotzé E, Van Huyssteen CW (2019) Soil, agriculture and food. In: Knight J,
Rogerson CM (eds) The geography of South Africa – contemporary changes and new
directions. Springer, Berlin, pp 111–124
2. Knight J, Holmes PJ (2018) Southern African drylands and their unique challenges. In:
Holmes PJ, Boardman J (eds) Southern African landscapes and environmental change.
Routledge, New York, pp 136–152
3. Knight J (2019) The making of the South African landscape. In: Knight J, Rogerson CM (eds)
The geography of South Africa – contemporary changes and new directions. Springer, Berlin,
pp 7–14
4. Department of Agriculture, Land Reform & Rural Development (2020) Abstract of agricul-
tural statistics 2020. Department of Agriculture, Land Reform & Rural Development, Pretoria
5. Van Huyssteen CW, Du Preez CC, Holmes PJ (2019) Agriculture and a changing biophysical
environment. In: Holmes PJ, Boardman J (eds) Southern African landscapes and environmen-
tal change. Routledge, New York, pp 228–248
6. Maluleke R (2020) Census of commercial agriculture (2017). Financial and production
statistics. Report no. 11–02-01 (2017). Statistics South Africa, Pretoria. http://www.statssa.
gov.za/publications/Report-11-02-01/Report-11-02-012017.pdf. Accessed 8 Feb 2021
7. Schulze RE, Maharaj M, Lynch SD, Howe BJ, Melvil-Thompson B (2001) South African atlas
of agrohydrology and climatology, Beta 1.002. University of KwaZulu-Natal,
Pietermaritzburg
8. Land Type Survey Staff (2000) Land types of South Africa. Agricultural Research Council –
Institute for Soil, Climate and Water, Pretoria
9. Fey MV (2010) Soils of South Africa: their distribution, properties, classification, genesis, use
and environmental significance. Cambridge University Press, New York
10. Soil Classification Working Group (1991) Soil classification – a taxonomic system for
South Africa. Memoirs on the agricultural natural resources of South Africa no 15. Department
of Agricultural Development, Pretoria
11. Pole Evans IB (1936) A vegetation map of South Africa. Memoirs of the botanical survey of
South Africa, vol 15, pp 1–23
12. Acocks JPH (1953) Veld types of South Africa. Memoirs of the botanical survey of
South Africa, vol 28, pp 1–192
13. Mucina L, Rutherford MC (2006) The vegetation of South Africa, Lesotho and Swaziland.
Strelitzia 19. South African National Biodiversity Institute, Pretoria. 807 pp
14. Finch JM, Meadows ME (2019) South African biomes and their changes over time. In:
Knight J, Rogerson CM (eds) The geography of South Africa – contemporary changes and
new directions. Springer, Berlin, pp 57–69
15. Agricultural Research Council – Institute for Soil, Climate and Water (2003) NDVI: long term
annual average. Agricultural Research Council – Institute for Soil, Climate and Water, Pretoria
16. Department of Agriculture, Forestry and Fisheries (2016) 2016 grazing capacity map of
South Africa. Department of Agriculture, Forestry and Fisheries, Pretoria
356 C. W. van Huyssteen and C. C. du Preez

17. Kotzé E, Snyman HA, Du Preez CC (2020) Rangeland management and soil quality in
South Africa. In: Lal R, Stewart BA (eds) Advances in soil science: soil degradation in Africa.
CRC Press, Boca Raton
18. Snyman HA (1999) Soil erosion and conservation. In: Tainton NM (ed) Veld management in
South Africa. University of Natal Press, Pietermaritzburg
19. Linstädter A, Kuhn A, Naumann C, Rasch S, Sandhage-Hofmann A, Amelung W, Jordaan J,
Du Preez CC, Bollig M (2016) Assessing the resilience of a real-world social-ecological
system: lessons from a multidisciplinary evaluation of a South African pastoral system. Ecol
Soc 21(3)
20. World Wildlife Fund (2010) Agriculture: facts and trends in South Africa. World Wildlife
Fund South Africa, Cape Town
21. Department of Agriculture, Forestry and Fisheries (2016) Abstract of agricultural statistics.
Department of Agriculture, Forestry and Fisheries, Pretoria
22. Doran JW, Saffley M (1997) Defining and assessing soil health and sustainable
productivity. In: Pankhurst C, Dube BM, Gupta VVSR (eds) Biological indicators of soil
health. CAB International, Wallingford
23. Lal R, Stewart BA (1992) Need for land restoration. In: Lal R, Stewart BA (eds) Soil
restoration. Advances in soil science. Springer, Berlin
24. Du Preez CC, Van Huyssteen CW (2020) Threats to soil and water resources in South Africa.
Environ Res 183:109015
25. Le Roux PAL, Du Preez CC (2006) Nature and distribution of South African plinthic soils:
conditions for soft and hard plinthic soils. S Afr J Plant Soil 23:120–126
26. Laker MC, Nortje GP (2019) Review of existing knowledge on soil crusting in South Africa.
Adv Agron 155:189–242
27. Agassi M, Shainberg I, Morin J (1981) Effect of electrolyte concentration and soil sodicity on
infiltration rate and crust formation. Soil Sci Soc Am J 45:848–851
28. der Watt V, HvH VC (1992) Soil crusting – the African view. In: Summer ME, Stewart BA
(eds) Soil crusting, chemical and physical processes. CRC Press, Boca Raton
29. Mills AJ, Fey MV (2004) Declining soil quality in South Africa: effects of land use on soil
organic matter and surface crusting. S Afr J Plant Soil 21:388–398
30. Fey MV, Mills AJ (2003) Declining soil quality in South Africa: effects of land use on soil
organic matter and surface crusting. S Afr J Sci 99:429–436
31. Menon M, Yuan Q, Dougill AJ, Hoon SR, Thomas AD, Williams RA (2011) Assessment of
physical and hydrological profiles of biological soil crusts using microtomography and
modelling. J Hydrol 397:47–54
32. Laker MC, Nortje GP (2020) Review of existing knowledge in subsurface soil compaction in
South Africa. Adv Agron 162:143–197
33. Du Preez CC, Bennie ATP, du Burger R (1981) Effect of implement traffic and irrigation on
soil compaction at Vaalharts. Agrochemophysica 13:7–12
34. Bennie ATP, Krynauw GN (1985) Causes, adverse effects and control of soil compaction. S
Afr J Plant Soil 2:109–114
35. Snyman HA, Du Preez CC (2005) Rangeland degradation in a semi-arid South Africa – II:
influence on soil quality. J Arid Environ 60:483–507
36. Nortje GP, Van Hoven W, Laker MC, Jodaan JC, Louw MA (2016) Quantifying the impact of
off-road driving on root-area distribution in soils. Afr J Wildl Res 46:33–48
37. Smith CW, Johnston MA, Lorentz SA (1997) Assessing the compactibility of South African
forestry soils. 1. The effect of soil type, water content and applied pressure on uni-axial
compaction. Soil Tillage Res 41:53–73
38. Du Preez CC, Van Huyssteen CW, Amehung W (2020) Changes in soil organic matter content
and quality in South African arable land. In: Lal R, Stewart BA (eds) Advances in soil science:
soil degradation in Africa. CRC Press, Boca Raton
Agricultural Land Degradation in South Africa 357

39. Lobe I, Sandhage-Hofmann A, Brodowski S, Du Preez CC, Amelung W (2011) Aggregate


dynamics and associated soil organic matter contents as influenced by prolonged arable
cropping in the South African Highveld. Geoderma 162:251–259
40. Kögel-Kraber I, Amelung W (2014) Dynamics, chemistry and preservation of organic matter
in soils. In: Holland HD, Furekian KK (eds) Treatise on geochemistry. Elsevier, Oxford
41. Kösters R, Du Preez CC, Amelung W (2013) Re-aggregation of degraded cropland soils with
prolonged secondary pasture management in the South African Highveld. Geoderma 192:173–
181
42. Scotney DM (1978) Advances in soil conservation and land use planning in Southern Africa,
1953–1978. Technical communication, vol 165. Department of Agriculture and Technical
Services, Pretoria
43. Laker MC (2004) Advances in soil erosion, soil conservation, land suitability evaluation and
land use planning research in South Africa, 1978–2003. S Afr J Plant Soil 21:345–368
44. Du Plessis MCF (1986) Grond agteruitgang. [Soil degradation]. S Afr J Nat Sci Technol 5:
126–168
45. Scotney DM, McPhee PJ (1992) Soil erosion and conservation. In: Van Oudshoorn FP
(ed) Guide to grasses of South Africa. Braza Publishers, Pretoria
46. Schoeman JL, Kock FG, Kaempffer LC, Scotney DM (1992) Wind erosion sensitive areas in
South Africa. Papers of 17th Soil Science Society of South Africa Congress, Stellenbosch
47. Van der Westhuizen AJ (1986) The influence of tillage practices on wind erosion in fields. M.
Sc. Agric. dissertation. University of the Free State, Bloemfontein
48. Snyman HA, Van Rensburg WJL (1986) Effect of slope and plant cover on run-off, soil loss
and water use efficiency of natural veld. J Grassland Soc S Afr 3:153–158
49. Garland GG (1990) Technique for assessing erosion risk from mountain footpaths. Environ
Manag 14:793–798
50. Arbuthnot FD (1995) Report of the ESA working group on land degradation. Directorate of
Resource Conservation, Department of Agriculture, Pretoria
51. Theron CHB (1988) The protection of steep lands under pineapples from erosion caused by
water. In: Proceedings 5th ISCO conference, Bangkok
52. Hill BI (1990) Cut soil losses with “plok” on new pineapple plantings. Dohne Agric 12:20–22
53. Grant PM (1978) Advances in soil fertility and crop nutrition research in Southern Africa
(1953–1978). Technical communication, vol 165. Department of Agriculture and Technical
Services, Pretoria
54. Barnard RO, Du Preez CC (2004) Soil fertility in South Africa: the last twenty-five years. S
Afr J Plant Soil 21:301–315
55. Biesenbach FW (1984) A NPK balance sheet for the agriculture soils of the Republic of
South Africa. Technical communication, vol 187. Department of Agriculture and Technical
Services, Pretoria
56. Skeen JB (1997) President’s report. Fertil Soc S Afr J:3–10. FSSA, Lynnwood Ridge
57. Brandt HC (2012) President’s report. Fertil Soc S Afr J:3–11. FSSA, Lynnwood Ridge
58. Scotney DM, Dijkhuis FJ (1990) Changes in the fertility status of South African soils. S Afr J
Sci 86:395–402
59. Du Preez CC (2003) Volhoubare landgebruik en grondkwaliteit: organiese materiaal as ‘n
indikator. [Sustainable land use and soil quality: organic matter as an indicator]. S Afr J Nat
Sci Technol 22:106–112
60. Du Preez CC, Steyn JT, Kotzé E (2001) Long-term effects of wheat residue management on
some fertility indicators of a semi-arid Plinthosol. Soil Tillage Res 63:25–33
61. Kotzé E, Du Preez CC (2008) Influence of long-term wheat residue management on acidity
and macronutrients in an Avalon soil. S Afr J Plant Soil 25:14–21
62. Loke PF, Kotzé E, Du Preez CC (2013) Impact of long-term wheat production management
practices on soil acidity, phosphorus and some micronutrients in semi-arid Plinthosol. Soil Res
51:415–426
358 C. W. van Huyssteen and C. C. du Preez

63. Wiltshire GH, Du Preez CC (1993) Long-term effects of conservation practices on the nitrogen
fertility of a soil cropped annually to wheat. S Afr J Plant Soil 10:70–76
64. Thierfelder C, Rusinamhodzi L, Ngwira AR, Mupangwa W, Nyagumbo I, Kassie GT, Cairns
JE (2015) Conservation agriculture in Southern Africa: advances in knowledge. Renew Agric
Food Syste 30:328–348
65. Sithole NJ, Magwaza LS, Matongoya PL (2016) Conservation agriculture and its impact on
soil quality and maize yield: a South African perspective. Soil Tillage Res 162:55–67
66. Fertilizer Association of Southern Africa (2017) Fertilizer handbook. Fertilizer Association of
Southern Africa, Lynnwood Ridge
67. Johnston MA (1991) Measurement of soil salinity using the four electrode probe. In: Pro-
ceedings South Africa irrigation symposium, Durban
68. Schmidt CJJ, Adriaanse FG, Du Preez CC (2007) Extractable soil phosphorus threshold values
for dryland maize in the South African Highveld. S Afr J Plant Soil 24:37–46
69. Venter AE, Du Preez CC (2021) Phosphorus extraction by selected methods in alkaline and
calcareous soils after mono-ammonium phosphate application. S Afr J Plant Soil 38:60–69
70. White VG, Hardie AG, Raath PJ (2020) Relationships between commonly used South African
and international soil phosphorus extraction tests in pristine and cultivated soils. S Afr J Plant
Soil 37:265–272
71. Ramaru JM, Hagmann J, Mamabolo ZM, Netshivhodza MH (2009) Innovation through
action – an action research journey with smallholder farmers in Limpopo Province,
South Africa: experiences of soil fertility management. In: Almekinders C, Beukeman L,
Tromp C (eds) Research in action – theories and practices for innovation and social change.
Wageningen Academic Publishers, Wageningen
72. Roberts VG, Adey S, Manson AD (2003) An investigation into soil fertility in two resource-
poor farming communities in KwaZulu-Natal (South Africa). S Afr J Plant Soil 20:146–151
73. Mkhabela TS, Materechera SA (2003) Factors influencing the utilization of cattle and chicken
manure for soil fertility management by emerging farmers in the moist Midlands of KwaZulu-
Natal Province, South Africa. Nutr Cycl Agroecosyst 65:151–165
74. Materechera SA (2010) Utilization and management practices of animal manure for
replenishing soil fertility among smallscale crop farmers in semi-arid farming districts of the
North West Province, South Africa. Nutr Cycl Agrecosyst 87:415–428
75. Scholes RJ (1990) The influence of soil fertility on the ecology of Southern African dry
savannas. J Biogeogr 17:415–419
76. Richards MB, Stock WD, Cowling RM (1997) Soil nutrient dynamics and community
boundaries in the fynbos vegetation of South Africa. Plant Ecol 130:143–153
77. Kotzé E, Sandhage-Hofmann A, Meinel JA, Du Preez CC, Amelung W (2013) Rangeland
management impacts on the properties of clayey soils along grazing gradients in the semi-arid
grassland biome of South Africa. J Arid Environ 97:220–229
78. Sandhage-Hofmann A, Kotzé E, Van Delden L, Dominiak M, Fouche HJ, Van der Westhuizen
HC, Oomen RJ, Du Preez CC, Amelung W (2015) Rangeland management effects on soil
properties in the savanna biome, South Africa: a case study along grazing gradients in
communal and commercial farms. J Arid Environ 120:14–25
79. Swanepoel PA, Botha PR, Du Preez CC, Snyman HA, Labuschagne J (2015) Managing
cultivated pastures for improving soil quality in South Africa: challenges and opportunities.
Afr J Range For Sci 32:91–96
80. Swanepoel PA, Du Preez CC, Botha PR, Snyman HA (2015) A critical view on the soil
fertility status of minimum till kikuyu–ryegrass pastures in South Africa. Afr J Range For Sci
32:113–124
81. Lechmere-Oertel RG, Cowling RM, Kerley GIH (2015) Landscape dysfunction and reduced
spatial heterogeneity in soil resources and fertility in semi-arid succulent, South Africa. Austral
Ecol 30:615–624
82. Blackmore AC, Mentis MT, Scholes RJ (1996) The origin and extent of nutrient-enriched
patches within a nutrient-poor savanna in South Africa. J Biogeogr 17:463–470
Agricultural Land Degradation in South Africa 359

83. Laker MC, Hensley M, Beyers CP, Van Rensburg SJ (1980) Environmental associations with
oesophageal cancer – an integrated model. S Afr Cancer Bull 24:69–70
84. Ceruti PO, Fey MV, Pooley J (2003) Soil nutrient deficiencies in an area of endemic
osteoarthritis (Mseni Joint Disease) and dwarfism in Maputuland, South Africa. In: Skinner
HCW, Berger AR (eds) Geology and health: closing the gap. Oxford University Press,
New York
85. ACC/SCN (2000) Fourth report on the world nutritional situation. ACC/SCN in Collaboration
with IFPRI, Geneva
86. Steyn CE, Herselman JE (2001) The status of soil quality in terms of trace elements in the
Mpumalanga Province. Report no GW/A/2001/19. Agricultural Research Council – Institute
for Soil, Climate and Water, Pretoria
87. Herselman JE, Steyn CE (2001) Predicted concentrations of trace elements in South African
soils. Report GW/A/2001/19. Agricultural Research Council – Institute for Soil, Climate and
Water, Pretoria
88. Beukes DJ (1995) Benefits from identifying and correcting soil acidity in agriculture. Agri-
cultural Research Council – Institute for Soil, Climate and Water, Pretoria
89. Beukes DJ (1997) Perspective on the soil acidity problem in South Africa. In: Proceedings of
soil acidity initiative symposium, Ermelo
90. Farina MPW (1997) Subsoil acidity and its management in South Africa. In: Proceedings of
soil acidity initiative symposium, Ermelo
91. Haumann PE (1997) Soil acidity management: essential condition for sustainable crop
production. In: Proceedings of soil acidity initiative symposium, Ermelo
92. Fey MV (2001) Consequences of a landscape turned over: the effect of excessive soil acidity
on the natural resources, agriculture and forestry. In: Proceedings 5th international plant soil
interactions at low pH symposium, South Africa
93. Haynes RJ (2001) Improving nutrient use efficiency as a tool for the management of acid
soils. In: Proceedings 5th international plant soil interactions at low pH symposium,
South Africa
94. Van der Merwe AJ, De Villiers MC, Barnard RO, Beukes DJ, Laker MC, Berry WAJ (2000)
Management and rehabilitation of acid and impoverished soils in South Africa. In: Proceed-
ings of FAO/ISCW expert consultation on management of degraded soils in Southern and East
Africa, Pretoria
95. Schoeman JL, Van Deventer PW (2004) Soils and the environment: the past 25 years. S Afr J
Plant Soil 21:369–387
96. Streutker A (1989) The role of irrigation farming, irrigation planner and irrigation manager in
food production and soil degradation. Abstracts of the combined Congress, Wild Coast,
Transkei
97. Nell JP, Van Huyssteen CW (2018) Prediction of primary salinity, sodicity and alkalinity in
South African soils. S Afr J Plant Soil 35:173–178
98. Du Plessis HM (1991) Researching and applying measures to conserve natural resources. In:
Proceedings South Africa irrigation symposium, Durban
99. Du Preez CC, Strydom MG, Le Roux PAL, Pretorius JP, Van Rensburg LD, Bennie ATP
(2000) Effects of water quality on irrigation farming along the lower Vaal River: the influence
on soils and crops. Report no. 740/1/00. Water Research Commission, Pretoria
100. Le Roux PAL, Du Preez CC, Srydom MG, Van Rensburg LD, Bennie ATP (2007) Effect of
irrigation on soil salinity profiles along the lower Vaal River, South Africa. Water SA 33:473–
478
101. Van Rensburg LD, Strydom MG, Du Preez CC, Bennie ATP, Le Roux PAL, Pretorius JP
(2008) Prediction of salt balances in irrigated soil along the lower Vaal River, South Africa.
Water SA 34:11–17
102. Van Rensburg LD, De Clercq WP, Barnard JH, Du Preez CC (2011) Salinity guidelines for
irrigation: case studies from Water Research Commission projects along the lower Vaal, Riet,
Berg and Breede Rivers. Water SA 37:739–749
360 C. W. van Huyssteen and C. C. du Preez

103. De Clercq WP, Fey MV, Moolman JH, Wessels WPJ, Eigenhuis B, Hoffman JE (2001)
Experimental irrigation of vineyards with saline water. Report no. 695/1/01. Water Research
Commission, Pretoria
104. Annandale JG, Du Preez CC (2005) Nutrients: agricultural contribution management and
modelling. Discussion paper 1. In: Rossouw JN, AHM G (eds) Knowledge review of model-
ling non-point source pollution in agriculture from field to catchment scale. Report no. 1467/1/
05. Water Research Commission, Pretoria
105. Van Niekerk AM (1992) Development of water quality management plan for the upper
Oliphant’s River Basin. In: Water week conference. Council for Scientific and Industrial
Research, Pretoria
106. Batchelor CR (1992) From damned to dammed. Fauna Flora 48:22–29
107. Clark JHA, Tredoux M, Van Huyssteen CW (2015) Heavy metals in the soils of Bloemfontein,
South Africa: concentration levels and possible sources. Environ Monit Assess 187. https://
doi.org/10.1007/s10661-015-4608-1
108. Barnard RO (2000) Carbon sequestration in South Africa soils. Report no. GW/A/2000/48.
Agricultural Research Council – Institute for Soil, Climate and Water, Pretoria
109. Rantoa NR, Van Huyssteen CW, Du Preez CC (2015) Organic carbon content in the soil
master horizons of South Africa. Vadose Zone J 14. https://doi.org/10.2136/vzj2014.10.0143
110. Soil Classification Working Group (2018) Soil classification – a natural and anthropogenic
system for South Africa. Agricultural Research Council Institute for Soil, Climate and Water,
Pretoria
111. Swanepoel CM, Rotter RP, Van der Laan M, Annandale JG, Beukes DJ, Du Preez CC,
Swanepoel LH, Van der Merwe A, Hoffmann MP (2018) The benefits of conservation
agriculture on soil organic carbon and yield in southern Africa are site-specific. Soil Tillage
Res 183:72–82
112. Dominy CS, Haynes RJ, Van Antwerpen R (2002) Loss of soil organic matter and related soil
properties under long-term sugarcane production on two contrasting soils. Biol Fertil Soils 36:
350–356
113. Van Antwerpen R, Meyer JH (1996) Soil degradation under sugarcane cultivation in northern
KwaZulu-Natal. Proceed S Afr Sugar Technol Assoc 70:29–33
114. Mthimkhulu S, Podwojewski P, Hughes J, Titshall L, Van Antwerpen R (2016) The effect of
72 years of sugarcane residues and fertilizer management on soil physico-chemical properties.
Agric Ecosyst Environ 225:54–61
115. Swanepoel PA, Habig J, Du Preez CC, Botha PR, Snyman HA (2014) Biological quality of a
podzolic soil after 19 years of irrigated minimum-till kikuyu–ryegrass pasture. Soil Res 52:64–
75
116. Swanepoel PA, Du Preez CC, Botha PR, Snyman HA, Habig J (2015) Assessment of tillage
effects on soil quality of pastures in South Africa using indexing methods. Soil Res 53:274–
285
117. Lebenya RM, Van Huyssteen CW, Du Preez CC (2018) Change in soil organic carbon and
nitrogen stocks eight years after conversion of sub-humid grassland to Pinus and Eucalyptus
forestry. Soil Res 56:318–330
118. Du Preez CC, Snyman HA (1993) Organic matter content of a soil in a semi-arid climate with
three long-standing veld conditions. Afr J Range For Sci 10:108–110
119. Haynes RJ, Graham MH (2004) Soil biology and biochemistry – a new direction for
South African soil science. S Afr J Plant Soil 23:320–344
120. Haynes RJ, Dominy CS, Graham MH (2003) Effect of agricultural land use on soil organic
matter status and the composition of earthworm communities in KwaZulu-Natal, South Africa.
Agric Ecosyst Environ 95:453–464
121. Graham MH, Haynes RJ, Meyer JH (2002) Soil organic matter content and quality: effects of
fertilizer applications, burning and trash retention on a long-term sugarcane experiment in
South Africa. Soil Biol Biochem 34:93–102
Agricultural Land Degradation in South Africa 361

122. Belay A, Claassens AS, Wehner RC (2002) Soil nutrient contents, microbial properties and
maize yield under long-term legume-based crop rotation and fertilization: a comparison of
residual effect of manure and N, P, K fertilizers. S Afr J Plant Soil 19:104–110
123. Belay A, Claassens AS, Wehner RC, De Beer JM (2001) Influence of residual manure on
selected nutrient elements and microbial composition of soil under long-term crop rotation. S
Afr J Plant Soil 18:1–6
124. Milne R, Haynes RJ (2004) Soil organic matter, microbial properties, and aggregate stability
under annual and perennial pastures. Biol Fertil Soils 39:172–178
125. Fynn RWS, Haynes RJ, O’Connor JGO (2003) Burning causes long-term changes in soil
organic matter content of a South African grassland. Soil Biol Biochem 35:577–687
126. Kotze E, Sandhage-Hofmann A, Amelung W, Oomen RJ, Du Preez CC (2017) Soil microbial
communities in different rangeland management systems of a sandy savannah and clayey
grassland ecosystem, South Africa. Nutr Cycl Agroecosyst 107:227–245
127. Dlamini JC, Haynes RJ, Van Antwerpen R (2001) Exotic earthworm species dominant in soils
on sugarcane estates in the Eshowe area of the north coast of KwaZulu-Natal. Proceed S Afr
Sugar Technol Assoc 75:217–221
128. Standing D, Killham K (2013) Sustainable management of soil and plant health by optimizing
soil biological function. In: Gregory PJ, Nortcliff S (eds) Soil conditions and plant growth.
Wiley – Blackwell, Chichester
129. Hoffman MT, Todd SW, Ntshona ZN, Turner SD (1999) Land degradation in South Africa.
National Botanical Institute, Cape Town
Agricultural Land Degradation
in the United States of America

Eric C. Brevik

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
2 Soil Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
3 Loss of Soil Organic Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
4 Overgrazing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
5 Salinization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
6 Acidification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
7 Soil Contamination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
8 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388

Abstract Soil management for agricultural purposes has been practiced in the
territory that makes up the United States of America (USA) for thousands of
years. However, the introduction of European agricultural management techniques
during colonization led to large increases in soil degradation. The greatest soil
management problems in modern USA agriculture are soil erosion and loss of soil
organic matter (SOM). Soil erosion rates have been reduced since the 1980s, but still
exceed sustainable levels in many parts of the country. Several management prac-
tices have been proposed to restore SOM levels, but more research is needed to really
understand their effectiveness. Other soil degradation issues in the USA include
overgrazing, salinization, acidification, and soil contamination. Rangelands and
pastures are about half of the land area in the USA, with a large portion of these at
least lightly overgrazed. Salinization is a concern with irrigation in dryland regions
and in semi-arid regions where soil–water relationships have been altered. Mainte-
nance of appropriate soil–water levels is critical in these situations. Acidification has
become a problem due to (a) acid rain, (b) long-term fertilizer use, and (c) the

E. C. Brevik (✉)
College of Agricultural, Life, and Physical Sciences, Southern Illinois University, Carbondale,
IL, USA
e-mail: [email protected]

Paulo Pereira, Miriam Muñoz-Rojas, Igor Bogunovic, and Wenwu Zhao (eds.), 363
Impact of Agriculture on Soil Degradation I: Perspectives from Africa, Asia, America
and Oceania, Hdb Env Chem (2023) 120: 363–392, DOI 10.1007/698_2022_918,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022,
Published online: 11 December 2022
364 E. C. Brevik

oxidation of soils that contain sulfides. Atmospheric emissions standards, appropri-


ate fertilizer management, and maintaining sulfide-containing materials under anaer-
obic conditions can help address these challenges. Soil contamination with trace
elements and organic chemicals has occurred due to fertilizer and pesticide use.
These issues are not unique to the USA, but this chapter presents them from a USA
perspective.

Keywords Acidification, Organic matter loss, Overgrazing, Salinization, Soil


erosion

1 Introduction

Agriculture, and presumably some level of soil management in support of that


agriculture, in what is now the United States of America (USA) dates to at least
7,000 years before present (BP) [1]. Intercropping, addition of ash to agricultural
soils, and fallowing were commonly used to maintain fertility [2] and crops were
planted in small mounds, which provided resistance to erosion [1] (Fig. 1). Native
American fields were also frequently located in landscape positions that benefited
their agriculture, providing advantages such as fertile soils found on floodplains or
soil properties that maximized root zone water content in drylands [3]. Water
harvesting techniques, rock mulch and irrigation were utilized in arid locations
[4, 5] and some tribes had systems of soil classification or took advantage of native

Fig. 1 A traditional Wampanoag garden at Plimoth Plantation, Massachusetts. Note the


intercropped maize and squash and abundant ground cover that reduced erosion. Although difficult
to see in this photograph, each maize plant was seeded into a small mound along with the squash.
Photograph by Eric Brevik
Agricultural Land Degradation in the United States of America 365

plants to identify areas of productive soils [2, 6]. One major soil management
technique missing from Native American agriculture was amending of soils with
animal manure, as North American tribes did not have domesticated livestock
[7]. The Native American approaches to agriculture were very attuned to nature,
and in some cases led to farming of the same fields for over 3,000 years with no
indications of soil degradation or decline in crop productivity [5]. However, due to a
lack of written records our knowledge of Native American agriculture is largely
restricted to archeological and anthropological studies and is therefore
incomplete [8].
Europe’s inability to feed itself was a major factor that led to a search for new
agricultural lands and launched European colonization, including immigration to the
lands that would become the modern USA [7]. The monoculture, plow-based, row
cropping methods of plant agriculture imported from Europe left soils more suscep-
tible to erosion and other forms of degradation than the agricultural methods of the
Native Americans [3]. As European farmers moved to the New World, they also
began to remove the forests from and farm steep hillsides [7]. This led to degradation
of soils and decreased crop productivity [9], sometimes in rather spectacular fashion
(Fig. 2). Not all Americans saw soil degradation as a necessary result of agriculture.

Fig. 2 Students from


Valdosta State University
visit Providence Canyon in
west-central Georgia, USA.
Providence Canyon consists
of gullies up to 45 m deep
and covering over 400 ha
that were formed by poor
farming practices in the
1800s. Photograph courtesy
of Judy Grable
366 E. C. Brevik

Thomas Jefferson, who would become the third President of the USA following
adoption of the present Constitution, was a leader in soil conservation and restora-
tion. However, the typical response to soil degradation was to pack up and move on
to new, fresh lands, something also inherited from European practices [7]. After the
USA formally gained independence from Great Britain in 1783 with the signing of a
peace treaty that ended the American Revolutionary War, the new nation rapidly
expanded westward [10], bringing European agricultural practices with it [11]. The
Native American practice of choosing prime agricultural lands based on their
landscape position gradually came to an end, and large-scale degradation was
introduced across large areas of the USA.
This chapter will review some of the major agricultural soil degradation chal-
lenges that have faced the USA, with an emphasis on the period from about 1900 to
the present. This is a period that occurs, for the most part, after the brief history given
above and more or less moves us into the era of modern agriculture within the USA,
where mechanical labor, widespread chemical treatments for nutrient supply and
pest control, and monoculture cropping came to dominate over traditional animal
labor and crop rotation systems. It will also explore some of the innovations that
have arisen to help address these challenges and ways that the situation might be
improved going forward.

2 Soil Erosion

Soil erosion represents a major threat to sustainable soil management [12]. By the
early 1900s that threat was gaining attention within the USA. William John McGee
and Edward Elway Free, both employees of the US Bureau of Soils, wrote extensive
reviews on what was known at the time regarding water and wind erosion, respec-
tively [13, 14]. Free’s (1911) publication is particularly noteworthy in that it appears
to be the first work to investigate aeolian processes from the perspective of their
impact on the soil resource, rather than as a geomorphic agent [15]. Early controlled
studies in soil erosion by water were also being set up in USA states like Ohio and
Missouri in the 1910s [16]. By the late 1920s Hugh Hammond Bennett, who would
go on to be recognized as the “Father of Soil Conservation” in the USA [17], was
loudly sounding the soil erosion alarm with publications extolling on the high cost of
soil erosion, which Bennett estimated cost American farmers $200 million year-1
(e.g., [18, 19]) (Fig. 3). Bennett’s efforts were instrumental in getting funding from
the US Congress to establish the Soil Erosion Service (SES) in 1933, and several
watershed-based experiment stations were rapidly established to start studying the
best ways to combat soil erosion [17] (Figs. 4 and 5).
The Dust Bowl, one of the greatest environmental disasters to hit modern
humanity, began in the 1930s. Choking clouds of dust blew across the Great Plains
region of the USA, burying and killing crops, livestock, and people [20] (Figs. 6 and
7a). By 1938 more than 5.4 million ha of land had lost at least 5 cm of topsoil
[21]. The Dust Bowl was most intense in the southern plains states of Colorado,
Agricultural Land Degradation in the United States of America 367

Fig. 3 H.H. Bennett standing beside a tree stump with roots exposed by erosion in Michigan USA,
ca. 1935. Evidence like this convinced Bennett that soil erosion was a national problem before it
was widely recognized as such in the USA. Photo courtesy of USDA-NRCS

Fig. 4 Planned soil erosion experiment stations in 1933, as shown by the colored dots. Map
courtesy of USDA-NRCS
368 E. C. Brevik

Fig. 5 A view of the Coon Valley watershed in Wisconsin, USA, in the 1930s. This was the first
soil erosion experimental station established by the Soil Erosion Service. Photo courtesy of the
USDA-NRCS

Fig. 6 The Great Plains region in the USA (shaded green). This region is particularly vulnerable to
land degradation due to its semi-arid environment
Agricultural Land Degradation in the United States of America 369

Fig. 7 (a) A dust storm blows into a Midwestern USA town during the 1930s. (b) A dust storm
blowing across the Lincoln Memorial in Washington, DC, in March 1935. Photos courtesy of the
USDA-NRCS

Kansas, New Mexico, Oklahoma, and Texas (Fig. 8), but it was felt in other parts of
the plains states as well. One of the author’s grandfathers grew up in western North
Dakota in the 1930s and told stories of drifts of dust to the roof of the house and
having to shovel dust to leave the house in the morning the way one might have to
shovel snow in the winter. Clouds of thick dust also made it all the way to major
cities along the east coast (Fig. 7b), and refuges from the Great Plains rolled into
California [7]. In short, no part of the USA was spared from the Dust Bowl’s effects,
and erosion became widely accepted as a problem in the USA.
Several steps were taken in response to the Dust Bowl. Bennett was convinced
that soil erosion was an agronomic problem and felt that there were agronomic
solutions [17]. Specifically, Bennett felt that erosion could be controlled using
vegetative cover while providing an economically viable option for farmers
[22]. Therefore, the watershed experiment stations focused on techniques such as
reducing tillage, crop rotations, contour strip-cropping on moderate slopes and
terraces with contour strips on relatively steep slopes, converting still steeper slopes
to pasture or hay, planting trees on the steepest slopes, and building fences to keep
out livestock. This approach reduced soil erosion by up to 75% on sloping agricul-
tural land [22]. Reduced tillage combined with various vegetative covers is often
referred to as conservation tillage. Wind erosion was addressed with standing
residues, wind barrier strips (taller crops planted in narrow rows between shorter
crops or fallowed fields), and tree rows (Fig. 9).
Today, many of the techniques developed in response to the erosion issues
highlighted by the Dust Bowl are still widely utilized in the USA, but it took
many decades to reach that point. In 2012 approximately 44% of USA croplands
were managed using some form of conservation tillage, but that percentage did not
rapidly grow until after the 1980s [23]. This growth has been effective at reducing
erosion. In the 1930s average soil erosion from croplands in the USA was about
16.3 t ha-1 [9, 24]. In 1989 soil erosion from croplands was similar, with an average
of about 16.0 t ha-1 (8.7 t ha-1 from water erosion and 7.3 t ha-1 from wind
370 E. C. Brevik

Fig. 8 A map of the areas most severely affected by the Dust Bowl between 1935 and 1938.
Recreation of a USDA Soil Conservation Service map from 1954, map is in the public domain
(https://en.wikipedia.org/wiki/Dust_Bowl#/media/File:Map_of_states_and_counties_affected_by_
the_Dust_Bowl,_sourced_from_US_federal_government_dept._(NRCS_SSRA-RAD).svg)

erosion). By 2017 this had been reduced to about 10.4 t ha-1 (6.0 t ha-1 water and
4.4 t ha-1 wind), with no meaningful change in erosion rates since 2007 [25]. How-
ever, rates did vary considerably across the country (Fig. 10). In particular, regions
that raise extensive row or small grain crops and have loess parent materials, such as
in Illinois, Iowa, Missouri, Nebraska, Tennessee, and Washington, had relatively
high rates of soil erosion.
One approach that has been used in the USA involves government programs that
pay farmers to take vulnerable lands out of production, an approach that is similar to
the current movement to pay farmers to manage their land in ways that provide
ecosystem services (e.g., [26, 27]). One popular such program was the Conservation
Reserve Program (CRP). Farmers enrolled in CRP received an annual payment to
plant sensitive land to species that would improve environmental health or quality,
with contracts lasting 10–15 years. Stated goals of the program included improved
water quality, reduced soil erosion, and providing wildlife habitat [28]. However,
Agricultural Land Degradation in the United States of America 371

Fig. 9 (a) Wind barrier strips in Montana, USA. (b) Tree rows in North Dakota, USA. Photos
courtesy of USDA-NRCS
372 E. C. Brevik

Fig. 10 Soil erosion by water on private lands in the USA in 1982 (top) and 2017 (bottom). Gray
through blue tones have erosion rates of 0–3.4 t ha-1, yellow and orange tones have erosion rates of
3.4–15.7 t ha-1, red tones have erosion rates of 15.7–22.4 t ha-1, and purple to black tones exceed
22.4 t ha-1. Map courtesy of US Department of Agriculture [25]
Agricultural Land Degradation in the United States of America 373

Fig. 11 Enrollment in the Conservation Reserve program, in millions of ha, by year. Data from US
Department of Agriculture [29]

government funding for CRP has declined in recent years and vulnerable acres are
coming out of the CRP contracts, with many of those acres being returned to
production agriculture (Fig. 11). CRP enrollment peaked at approximately 14.6
million ha in 2007 but declined to about 8.5 million ha in 2020 [29]. This has led
to considerable concern about loss of the environmental benefits CRP was designed
to provide, including reduced soil erosion. It is also important to note that semi-arid
areas are particularly susceptible to land degradation. The Great Plains region of the
USA is a semi-arid region and has been identified by Lepers et al. [30] as the portion
of the USA that may be the most susceptible to degradation, with both wind and
water erosion being of concern.

3 Loss of Soil Organic Matter

Soil organic matter (SOM) content is important for a number of soil physical,
chemical, and biological properties. Because of that, it is also important to crop
yield, with yields frequently being positively correlated to SOM content [31]. Soil
organic carbon (SOC) is often used as a proxy for SOM, and the levels of SOC vary
considerably across the USA (Fig. 12). Generally speaking, SOC content is expected
to be relatively high along the east coast, particularly in the northeastern part of the
USA, and low in the southwest. This is due to climate patterns, with high rainfall
along the east coast and cool temperatures in the northeast, while the southwestern
USA has a hot arid climate.
374 E. C. Brevik

Fig. 12 Average soil organic carbon stocks in various regions of the USA to a depth of 1 m
[32]. Image courtesy of USDA-NRCS

Fig. 13 The effects of soil


erosion on soil organic
matter (SOM) content in
Iowa soils. The soils, from
left to right, have been
slightly to severely eroded.
Note the thick dark layer on
top of the left soil sample,
indicating high SOM
content versus the lighter
brown colors of the middle
and right soil samples. Photo
courtesy of USDA-NRCS

Agricultural management has led to a decline in SOC in many soils [33] (Fig. 13).
Lal et al. [34] report that croplands in the USA lost 3–5 Pg of C between 1750 and
1950. This represents a loss of approximately 50% of the original, naturally occur-
ring SOC in these soils [35, 36] (Fig. 14). The loss of SOC is due to microbial
decomposition following the introduction of oxygen into the soil during tilling, loss
Agricultural Land Degradation in the United States of America 375

Fig. 14 Expected response of soil organic carbon (SOC) content to changes in management. The
soil is under a natural system from t0 to t1. At t1 the soil was managed for agriculture using a
traditional moldboard plow system, as was common in the USA through much of agricultural
history, with the red curve showing the expected decline in SOC before leveling out at a new
equilibrium, while the gray wavy line shows expected SOC if the soil had remained under a natural
system. At t2 the management changed to a no-till system (brown curve), which led to more SOC
than if conventional tillage had been maintained (straight gray line) but less than if the natural
system had remained in place (gray wavy line)

Fig. 15 Two views of strong granular structure in a topsoil. This structure is associated with
sufficient soil organic matter levels, good water infiltration and aeration, and easy root penetration.
Photos courtesy of USDA-NRCS

of the organic-rich surface soil layer to erosion, and lack of fresh organic additions to
the soil when large amounts of the crop are removed during harvest. Losses of SOM
cause a loss of soil structure (Fig. 15), which in turn causes decreased infiltration,
water storage, aeration and makes it more difficult for plant roots to penetrate the
soil, all of which affect crop yields [37]. The large and heavy equipment used in
modern agriculture also causes intense and deep soil compaction, which has effects
376 E. C. Brevik

Fig. 16 Intense compaction


from wheel traffic in the
interrow area (white arrow)
has restricted the root
growth from this maize
crop. Photo courtesy of the
USDA-NRCS

Fig. 17 The tractors typically used in the USA for field work in years past (left) were much smaller
and weighed considerably less than modern four-wheel-drive tractors (right). The extra weight
creates additional soil compaction concerns. Photos courtesy of USDA-NRCS

that are similar to those from structure loss; loss of structure is one effect of
compaction (Fig. 16). Tractors commonly used in the USA in the 1940s weighted
less than 2.7 metric tons, while today’s large four-wheel-drive tractors weigh about
18 metric tons [38] (Fig. 17).
A major point in the decline of SOM in the USA was likely the Homestead Act of
1862. This Act gave 65 ha of land owned by the US Government to anyone who was
21 or more years of age, a citizen of the USA or had stated plans to become one, paid
a $10 filing fee, and lived on and farmed the land for 5 years. In total, over
109 million ha of land was given out through the Homestead Act [39]. Because
Agricultural Land Degradation in the United States of America 377

the Homestead Act required working the land, and the moldboard plow was the
tillage tool of preference at the time, large areas of land that had been undisturbed
were plowed, initiating changes in soil properties. The decomposition of 1% of the
SOM in a typical natural soil would release about 1080 kg of N, 260 kg of P and
190 kg of S [35]. During the rush to acquire land following implementation of the
Homestead Act there was plenty of advertising that exaggerated the quality of the
available soils [40]. The nutrients released by decomposition of SOM can be taken
up by crops, and their release upon working of native soils with the moldboard plow
likely contributed to the stories of rich, fertile soils that helped bring new settlers
under the Homestead Act. However, as SOM levels decline nutrients in the soil are
depleted, and eventually the rush of natural fertilizer declines.
Soil scientists in the USA have long recognized loss of SOM as a serious concern.
Conservation tillage methods, including no-till, have been promoted by many as a
way to rebuild lost SOM in production agriculture fields (e.g., [36, 41–43]) (Fig. 14).
However, this has been questioned in recent years, with some studies now indicating
that SOC may be sequestered in the top few cm of soils being managed with
conservation tillage, but when the whole soil profile is considered there is no net
sequestration occurring (e.g., [44, 45]). In essence, these studies suggest only
sampling the top 10–30 cm of the soil is not sufficient to capture all of the SOC
dynamics that are occurring. VandenBygaart et al. [46] reviewed 62 Canadian
studies and concluded that no-till was most able to sequester SOC when average
annual precipitation was <550 mm year-1, suggesting that climate plays an impor-
tant role. Carr et al. [47] supported these findings, in that no-till sequestered SOC in
soils of western North Dakota, where average annual precipitation was about
400 mm. Whether or not cover crops, manures, and other management techniques
are included also makes a difference in the results obtained with any given tillage
system [37]. And finally, there is a lack of long-term studies in many areas, and
pedogenesis takes long periods of time. Therefore, more work in this area is needed
to determine the best practices to build SOM, and these practices may vary
depending on the environmental conditions of the locations in question.

4 Overgrazing

Rangelands and pastures represent approximately half of the land area in the USA
[48]. When properly managed, rangelands can provide meaningful food value
through livestock production. However, overgrazing can damage vegetative cover,
change forage species distribution including selecting for undesirable species,
increase erosion, decrease SOM, and cause a number of associated damages to soil
properties [49] (Fig. 18). The invasion of non-native grass species, sometimes
introduced during poorly conceived reclamation attempts, also increases the suscep-
tibility of rangeland to wildfires and climate change, and all of the above reduces the
ecosystem services that can be provided [50]. In the USA, overgrazed grasslands are
primarily found in the western part of the country. Of these, about 14.5% are lightly
378 E. C. Brevik

Fig. 18 Overgrazed pasture (left) beside well-managed pasture (right). Photo courtesy of USDA-
NRCS

overgrazed and 73.4% moderately overgrazed. The remaining 12.1% are strongly
overgrazed and are largely confined to the southwestern part of the country
[49]. Another concern in pastures and rangelands is degradation of stream banks
[9]. This can have adverse effects on both forage, as land is lost to the stream, and on
water quality, as sediment load increases. Stream banks can be protected by limiting
the areas livestock have access to and armoring the parts of the banks and stream bed
they come into contact with (Fig. 19).
The precise objectives when restoring degraded rangeland depend on the use that
the restored land will be put to. However, in all cases the primary reclamation goal is
to establish a desirable plant community that has low susceptibility to invasive
species and will support desired land use(s) [48].

5 Salinization

Salinization refers to the accumulation of water-soluble salts in a soil. It is both a


chemical and physical concern as relates to soil degradation [30]. Salinization is a
chemical concern in the way that it influences the ionic make-up of the soil solution
and resulting nutrient interactions. It is a physical concern in the way that it can cause
loss of soil structure and restrict plant access to water. Salinization is typically
caused by changes in a soil’s hydrology, which can be induced by irrigation
(Fig. 20) or changes in vegetative cover brought about by agriculture [51]. Because
irrigation is a common source of salinization, waterlogging is a common problem
Agricultural Land Degradation in the United States of America 379

Fig. 19 (a) A stream through a pasture that provides livestock unlimited access. Note the eroded
stream banks and exposed soil. (b) A managed stream crossing for livestock. Note the gravel cover
on the banks to prevent damage and erosion and the fencing that crosses the stream to limit livestock
access. Both pictures from Michigan, USA. Photos courtesy of USDA-NRCS
380 E. C. Brevik

Fig. 20 Salt accumulation (white on top of the ridges) due to capillary rise in a furrow-irrigated
system in the western USA. Note that the crop has been planted below the salts, but eventually salt
accumulation will reach a point where cropping will not be possible if soil remediation is not
undertaken. Photo courtesy of USDA-NRCS

that accompanies salinization. Approximately 23% of the irrigated land in the USA
faces the twin problems of salinization and waterlogging [52]. In the USA, the
regions of greatest vulnerability to salinization problems fall into the western portion
of the country, particularly in the southwestern and north-central regions.
Salinization problems are greatest in arid and semi-arid regions, where ground-
water often has a higher natural salt content than in more humid regions [51, 52]. It is
also common that water does not infiltrate all the way through the soil profile in these
dryland regions. Rather, evaporation tends to pull water up through the profile, leaving
dissolved salts behind as the water evaporates off the soil surface [53]. This process is
capable of accumulating large quantities of salts in a short period of time under
irrigated conditions (Fig. 20); up to 6.7 tons of salts has been documented ha-
1
year-1 in studies in the USA [51]. There are currently about 18.1 mha of irrigated
land in the USA, of which about 23% suffers from salinization [54].
Some areas of dryland farming in the USA are susceptible to the creation of
saline-seeps. These result from changes in vegetative cover leading to changes in the
hydrologic regime when the right (or wrong, depending on how one chooses to view
it) geological conditions are present. In summary, farming introduces a crop that
does not have a well-developed, living root system throughout the growing period.
This means the crop takes up less water than would be utilized by native vegetation,
and excess water infiltrates below the root zone, carrying dissolved salts with it. The
local geology includes an aquiclude below the root zone, so the infiltrating water
perches on this low permeability layer and runs along it down gradient. At some
point this aquiclude intersects the land surface, creating a seep that leaves salts
Agricultural Land Degradation in the United States of America 381

Fig. 21 Generalized diagram showing the creation of a saline seep. Figure courtesy of USDA-
NRCS [55]

behind as the water evaporates (Fig. 21). This process is particularly common in the
northern plains region of the USA, including Montana, North Dakota, and South
Dakota (Fig. 22).
Addressing salinization requires understanding the processes that drive it. In
irrigated systems, the problem is often too much water, which raises the local
water table and allows evaporation from the surface to wick water up through
capillary rise, leaving salts behind as the water evaporates. This is a particular
problem if the irrigated area is underlain by groundwater with a high dissolved
solids content [52]. Careful water application to avoid raising the water table [51]
and installing a drainage system [52] are techniques that have been used to address
this problem. For saline-seeps, the goal is to restore a more natural hydrology to the
agroecosystem, reducing the amount of water infiltrating through the soil profile,
which lowers the local water table and prevents the seepage of saline groundwater at
the saline seep site. This has been accomplished using alfalfa as a forage crop [56],
various cover crops [57], or by planting such sites to forages [58].
Saline soils can also be reclaimed by treating with gypsum and leaching with
high-quality water. However, this approach is very expensive and can only be used if
there is a source of high-quality water to leach salts from the soil with, along with a
way to drain leach waters away from the affected soil. This process also adds salts to
local water bodies, which is a drawback [59]. Other work on the remediation of salt-
contaminated soils has included the use of crystallization inhibitors to bring salts to
the soil surface, where they can be removed from the site [60].
382 E. C. Brevik

Fig. 22 A saline seep in South Dakota, USA. Photo courtesy of USDA-NRCS

6 Acidification

Soil acidification in the USA has been driven by (1) acid rain, (2) fertilizer use, and
(3) the oxidation of soils that contain sulfides (acid sulfate soils). Acid rain became a
major problem in the USA, particularly the northeastern region, in the 1900s
primarily due to emissions of sulfur and nitrogen-based gases from industrial
activities and the burning of fossil fuels [61]. The acidification of soils due to acid
rain led to the leaching of base cations, increases in exchangeable aluminum, and a
decrease in soil pH [62], all of which are negatives in most plant production. In 1970
Agricultural Land Degradation in the United States of America 383

the US Congress passed the Clean Air Act, which began to regulate emissions of a
variety of pollutants into the atmosphere [63]. The annual wet sulfate deposition
levels, a common measurement of acid rain, decreased 68% from 1989–1991 to
2017–2019 ([63]; Fig. 23). Corresponding recovery of soil properties has been
documented, although many soils still have a way to go before they return to their
pre-acid rain conditions [62].
Soil acidification due to the long-term use of nitrogen fertilizers has also been
reported in the USA. Negative effects on the soil include increases in exchangeable
acidity and decreases in cation exchange capacity, exchangeable calcium and mag-
nesium, and lower pH [64]. Even in regions of the USA where soil acidification has
not previously been considered a problem due to high natural soil pH levels, decades
of chemical nitrogen fertilizer use is leading to problems such as declining pH and
aluminum toxicity [65, 66]. The degree of acidification depends on the type of
nitrogen fertilizer used and the rate at which it is applied. Acidification is a particular
problem with ammonium-N, while applying only the amount of nitrogen needed and
nothing more minimizes acidification problems [67, 68]. Proper crop rotations are
also critical to reverse soil acidification from nitrogen fertilization [65].
The problem of acid sulfate soils does not cover as many ha as the previous two
issues, in fact, the USA is not considered one of the world’s major acid sulfate soil
regions [69]. However, these soils can be an important challenge locally. There are
three main places where this has been identified as a problem in the USA, (1) drained
coastal wetlands, (2) dredged marine sediments, and (3) mine spoils or other soils/
sediments that have been taken from anaerobic settings and deposited into an aerobic
setting [69]. Acid sulfate soils can form naturally or anthropogenically, but the
common theme is oxidation of sulfide minerals in previously anaerobic materials.
The pH in acid sulfate soils can drop from near neutral to less than 3 as large amounts
of sulfuric acid are created by the oxidation [70]. If this process occurs near a body of
water, aquatic ecosystems can be devastated by a flood of metals into the system
[71]. On land, the low pH and release of metals into the soil system can cause severe
loss of organismal abundance and diversity, including loss of crop yields, and
subsidence can become a problem as soils may lose more than 50% of their volume
[72]. Builders have been sued over the construction of developments on acid sulfate
soils that then caused problems for the owners [73]. The best management strategy
for acid sulfate soils is to leave them undisturbed. However, if disturbance is
necessary, these materials should be disturbed as little as is required for the work
to be done, and if possible, sulfide-containing materials should be restored to an
anaerobic setting once the work is completed [72].

7 Soil Contamination

Soils in the USA have been contaminated in several ways, including industrial
accidents, illegal dumping, and improper landfill design. Such contamination is
important from a human and environmental health perspective and is a major reason
384 E. C. Brevik

Fig. 23 (a) Average annual wet sulfate deposition in the USA, 1989–1991. (b) Average annual wet
sulfate deposition in the USA, 2017–2019. Maps courtesy of US EPA
Agricultural Land Degradation in the United States of America 385

for the environmental movement that led to passage of the first environmental laws
in the USA in the 1970s [74]. However, these sources of contamination are beyond
the scope of this chapter, which focuses on agricultural lands.
There have been instances of contamination of agricultural lands with trace
elements, organic chemicals, or the products of the mining and extraction industry.
Metals such as cadmium can be present in fertilizers as natural impurities, leading to
concern that contamination might occur given long-term applications. Studies to
investigate this seem to indicate this concern is negligible [75]. However, trace
elements have been introduced to fertilizers as fill materials in amounts that have
caused problems. In the 1980s and 1990s a company in the State of Washington was
blending hazardous wastes into their fertilizers as a way to get around following state
and federal hazardous waste regulations and paying the associated fees, as there were
no regulations on fertilizer fillers [76]. Basically, farmers were paying the company
to dispose of the company’s hazardous wastes, rather than the company paying to do
so. At about the same time, a widely used garden fertilizer in the State of Minnesota
was revealed to have high levels of arsenic, lead, and cadmium in it. In both cases,
the States passed laws (1998 in Washington, 2003 in Minnesota) to regulate the
composition of fertilizer fillers. Likewise, a number of trace elements have been used
in the past as insecticides, particularly in the fruit industry. This has left a number of
orchard soils contaminated with elements such as arsenic, and the banning or
regulation of many of these pesticide products [76].
Organic chemicals are widely used in modern agriculture as fungicides, herbi-
cides, and insecticides. Within the USA, approximately 20,000 people receive
hospital treatment each year for pesticide exposure, and about 30 of those die
[76]. Many organic chemicals are persistent in the soil environment, resisting
degradation by soil microorganisms [77]. This allows them to cause human and
environmental health problems for years after their original application. Some of
these chemicals cause so many problems in the environment they have become
known as persistent organic pollutants (POPs). There were originally 12 POPs,
known as the “Dirty Dozen,” placed on the Stockholm Convention list in 2001.
Another 17 have been added since, giving 29 POPs at present. Some of the POPs are
permanently banned internationally from use, but others have a limited range of
approved uses (Table 1). Best management practices with organic chemicals include
handling and applying these chemicals according to manufacturer’s instructions and
the use of personal protective equipment when handling them (Fig. 24).

8 Concluding Remarks

Soil degradation in the USA occurs for several reasons. The most widespread of
these are soil erosion and loss of soil organic matter. However, other issues such as
overgrazing, salinization, acidification, and soil contamination are also concerns that
must be considered and addressed. Many of these issues are not unique to the USA,
but this chapter has sought to explain them from the perspective of where in the USA
386 E. C. Brevik

Table 1 Persistent organic pollutants (POPs) identified by the Stockholm Convention (SC). Some
POPs can still be used for specific purposes as outlined in the SC. Table from Brevik et al. [78]
Year Annex
Chemical added Source in SC Additional notes
Aldrin 2001 P A
Chlordane 2001 P A
Dichlorodiphenyltrichloroethane 2001 P B DDT still used against mosqui-
(DDT) toes in several countries to con-
trol malaria
Dieldrin 2001 P A
Endrin 2001 P A
Heptachlor 2001 P A
Hexachlorobenzene (HCB) 2001 P, IC, A&C
UP
Mirex 2001 P A
Toxaphene 2001 P A
Polychlorinated biphenyls (PCB) 2001 IC, UP A&C Has specific exemptions under
Annex A
Polychlorinated dibenzo-p- 2001 UP C
dioxins (PCDD)
Polychlorinated dibenzofurans 2001 UP C
(PCDF)
Alpha hexachlorocyclohexane 2009 P A No exceptions or acceptable
uses
Beta hexachlorocyclohexane 2009 P A No exceptions or acceptable
uses
Chlordecone 2009 P A No exceptions or acceptable
uses
Hexabromobiphenyl 2009 IC A No exceptions or acceptable
uses
Hexabromodiphenyl ether, 2009 IC A Can be used in accordance with
heptabromodiphenyl ether the provisions of Part IV of
Annex A
Lindane 2009 P A Human use for control of head
lice and scabies as second-line
treatment
Pentachlorobenzene 2009 P, IC, A&C No exceptions or acceptable
UP uses
Perfluorooctane sulfonic acid, its 2009 IC A&B Acceptable purposes and spe-
salts and perfluorooctane sulfonyl cific exemptions in accordance
fluoride with Part III of Annex B,
amended 2019
Tetrabromodiphenyl ether, 2009 IC A Has specific exemptions under
pentabromodiphenyl ether Part V of Annex A
Technical endosulfan and its 2011 P A Exemptions for crop-pest com-
related isomers plexes in accordance with the
provisions of part VI of
Annex A
(continued)
Agricultural Land Degradation in the United States of America 387

Table 1 (continued)
Year Annex
Chemical added Source in SC Additional notes
Hexabromocyclododecane 2013 IC A Expanded and extruded poly-
styrene in buildings in accor-
dance with the provisions of part
VII of Annex A
Hexachlorobutadiene 2015 IC, UP A&C No exceptions or acceptable
uses, added to annex C in 2017
Pentachlorophenol and its salts 2015 P A Pentachlorophenol for utility
and esters poles and cross-arms in accor-
dance with the provisions of part
VIII of Annex A
Polychlorinated naphthalenes 2015 IC, UP A&C Can be used for production of
polyfluorinated naphthalenes,
including octafluoronaphthalene
Decabromodiphenyl ether 2017 IC A Exemptions for certain uses in
vehicles, aircraft, textiles, addi-
tives in plastic housings, etc.,
polyurethane foam for building
insulation
Short-chain chlorinated paraffins 2017 IC A Allowed as additives in trans-
mission belts, rubber conveyor
belts, leather, lubricant addi-
tives, tubes for outdoor decora-
tion bulbs, paints, adhesives,
metal processing, plasticizers
Dicofol 2019 P A No exceptions or acceptable
uses
P pesticide, IC industrial chemicals, UP unintentional production/by-products

Fig. 24 Proper personal protective equipment is important when handling organic chemicals in an
agricultural setting. Photo courtesy of USDA-NRCS
388 E. C. Brevik

they are most important and what has been done specifically within the USA to
address them. In most cases, the situation today is better than it was a few decades
ago. Soil erosion is down, the healthy soils movement is working to restore soil
organic matter, and the other issues discussed have generated considerable research
and efforts to reverse or slow them. However, there is still work to be done if the
USA is to achieve the zero net land degradation by 2030 goal set forth by the United
Nations Convention to Combat Desertification in 2012.

References

1. Warren DM (1994) Indigenous agricultural knowledge, technology, and social change. In:
McIsaac G, Edwards WR (eds) Sustainable agriculture in the American Midwest: lessons from
the past, prospects for the future. University of Illinois Press, Urbana, pp 35–53
2. Vazquez JM (2011) The role of indigenous knowledge and innovation in creating food
sovereignty in the Oneida Nation of Wisconsin. Unpublished Master’s thesis, Iowa State
University, Ames
3. Brevik EC, Fenton TE, Homburg JA (2016) Historical highlights in American soil science —
Prehistory to the 1970s. Catena 146:111–127
4. Homburg JA, Sandor JA (2011) Anthropogenic effects on soil quality of ancient agricultural
systems of the American Southwest. Catena 85:144–154
5. Homburg JA, Sandor JA, Norton JB (2005) Anthropogenic influences on Zuni agricultural
soils. Geoarchaeology: An International Journal 20:661–693
6. Pawluk RR (1995) Indigenous knowledge of soils and agriculture at Zuni Pueblo, New Mexico.
Unpublished master’s thesis, Departments of Agronomy and Anthropology, Iowa State
University, Ames
7. Montgomery DR (2007) Dirt: the erosion of civilizations. University of California Press,
Berkeley
8. Brevik EC, Homburg JA, Sandor JA (2018) Soils, climate, and ancient civilizations. In:
Horwath W, Kuzyakov Y (eds) Changing soil processes and ecosystem properties in the
Anthropocene. Developments in soil science series. Elsevier, Amsterdam, pp 1–28
9. Troeh FR, Hobbs JA, Donahue RL (2004) Soil and water conservation for productivity and
environmental protection.4th edn. Prentice Hall, Upper Saddle River
10. Miller RJ (2011) American Indians, the doctrine of discovery, and manifest destiny. Wyoming
Law Rev 11:329–349
11. McCoy K (2014) Manifesting destiny: a land education analysis of settler colonialism in
Jamestown, Virginia, USA. Environ Educ Res 20:82–97
12. Pimentel D, Harvey C, Resosudarmo P, Sinclair K, Kurz D, McNair M et al (1995) Environ-
mental and economic costs of soil erosion and conservation benefits. Science 267(5201):
1117–1123
13. Free EE (1911) The movement of soil material by the wind. U.S. Government Printing Office,
Washington DC. USDA Bureau of Soils Bulletin No. 68
14. McGee WJ (1911) Soil erosion. Bureau of Soils Bulletin No. 71. U.S. Government Printing
Office, Washington DC
15. Brevik EC (2004) Contributions of Edward Elway free to American soil science in the early
1900s. Soil Sci Soc Am J 68:904–906
16. Brevik EC (1999) George Nelson Coffey, Early American Pedologist. Sci Soc Am J 63:1485–
1493
17. Helms D (2010) Hugh Hammond Bennett and the creation of the Soil Conservation Service. J
Soil Water Conserv 65(2):37A–47A
Agricultural Land Degradation in the United States of America 389

18. Bennett HH (1928) The wasting heritage of the nation. Sci Mon 27(2):97–124
19. Bennett HH (1929) Uncle Sam Spendthrift—IX. Sci Am 136(4):237–239
20. Worster D (2004) Dust bowl: the southern plains in the 1930s. Oxford University Press,
New York
21. Hansen ZK, Libecap GD (2004) Small farms, externalities, and the Dust Bowl of the 1930s. J
Polit Econ 112(3):665–694
22. Anderson R (2002) Coon Valley days. Wisconsin Acad Rev 48(2):42–48
23. Duzy LM, Price AJ, Balkcom KS, Kornecki TS (2015) Conservation tillage under threat in the
United States. Outlooks Pest Manag 26:257–262
24. US Department of Commerce (1930) Farm acreage and farm values by townships or other
minor civil divisions. Bureau of the Census. United States Government Printing Office,
Washington, DC
25. US Department of Agriculture (2020) Summary Report: 2017 National Resources Inventory,
Natural Resources Conservation Service, Washington, DC, and Center for Survey Statistics and
Methodology, Iowa State University, Ames, Iowa. https://www.nrcs.usda.gov/wps/portal/nrcs/
main/national/technical/nra/nri/results. Accessed 31 July 2021
26. Goldstein JH, Presnall CK, López-Hoffman L, Nabhan GP, Knight RL, Ruyle GB, Toombs TP
(2011) Beef and beyond: paying for ecosystem services on western US rangelands. Rangelands
33:4–12
27. Gutman P (2007) Ecosystem services: foundations for a new rural–urban compact. Ecol Econ
62:383–387
28. Brevik EC, Steffan JJ, Burgess LC, Cerdà A (2017) Links between soil security and the
influence of soil on human health. In: Field D, Morgan C, McBratney A (eds) Global soil
security. Progress in soil science series. Springer, Rotterdam, pp 261–274
29. US Department of Agriculture (2020) The conservation reserve program: a 35 year history.
https://www.fsa.usda.gov/Assets/USDA-FSA-Public/usdafiles/Conservation/PDF/35_
YEARS_CRP_B.pdf. Accessed 31 July 2021
30. Lepers E, Lambin EF, Janetos AC, Defries R, Achard F, Ramankutty N, Scholes RJ (2005) A
synthesis of information on rapid land-cover change for the period 1981–2000. Bioscience 55:
115–124
31. Lal R (2020) Soil organic matter content and crop yield. J Soil Water Conserv 75(2):27A–32A
32. Wills S (2016) Weighted mean SOC stocks to 100 cm: region (MO) means. https://www.nrcs.
usda.gov/Internet/FSE_DOCUMENTS/nrcs142p2_052842.pdf. Accessed 1 Aug 2021
33. Brevik EC (2012) Soils and climate change: gas fluxes and soil processes. Soil Horizons 53(4):
12–23
34. Lal R, Follett RF, Kimble JM (2003) Achieving soil carbon sequestration in the United States: a
challenge to the policy makers. Soil Sci 168:827–845
35. Baumhardt RL, Stewart BA, Sainju UM (2015) North American soil degradation: processes,
practices, and mitigating strategies. Sustain For 7:2936–2960
36. Grace P, Robertson GP (2021) Soil carbon sequestration potential and the identification of
hotspots in the Eastern Corn Belt of the United States. Soil Sci Soc Am J. https://doi.org/10.
1002/saj2.20273
37. Brevik EC (2009) Soil health and productivity. In: Verheye W (ed) Soils, plant growth and crop
production. Encyclopedia of Life Support Systems (EOLSS), EOLSS Publishers, Oxford.
http://www.eolss.net. Accessed 1 Aug 2021
38. DeJong-Hughes J (2018) Soil compaction. University of Minnesota Extension. https://
extension.umn.edu/soil-management-and-health/soil-compaction. Accessed 1 Aug 2021
39. Potter LA, Schamel W (1997) The Homestead Act of 1862. Soc Educ 61:359–364
40. Anderson HL (2011) That settles it: the debate and consequences of the Homestead Act of 1862.
Hist Teach 45:117–137
41. Mowrer J, Rajan N, Sarangi D, Zapata D, Govindasamy P, Maity A, Singh V (2020) Soil
management practices of major crops in the United States and their potential for carbon
390 E. C. Brevik

sequestration. In: Ghosh P, Mahanta S, Mandal D, Mandal B, Ramakrishnan S (eds) Carbon


management in tropical and sub-tropical terrestrial systems. Springer, Singapore, pp 71–88
42. Novak JM, Watts DW, Bauer PJ, Karlen DL, Hunt PG, Mishra U (2020) Loamy sand soil
approaches organic carbon saturation after 37 years of conservation tillage. Agron J 112:3152–
3162
43. Sperow M (2016) Estimating carbon sequestration potential on U.S. agricultural topsoils. Soil
Tillage Res 155:390–400
44. Baker JM, Ochsner TE, Venterea RT, Griffis TJ (2007) Tillage and soil carbon sequestration-
what do we really know? Agric Ecosyst Environ 118:1–5
45. Christopher SF, Lal R, Mishra U (2009) Regional study of no-till effects on carbon sequestra-
tion in the Midwestern United States. Soil Sci Soc Am J 73:207–216
46. VandenBygaart AJ, Gregorich EG, Angers DA (2003) Influence of agricultural management on
soil organic carbon: a compendium and assessment of Canadian studies. Can J Soil Sci 83:363–
380
47. Carr PM, Brevik EC, Horsley RD, Martin GB (2015) Long-term no-tillage sequesters soil
organic carbon in cool semi-arid regions. Soil Horizons 56(6). https://doi.org/10.2136/sh15-
07-0016
48. DiTomaso JM, Masters RA, Peterson VF (2010) Rangeland invasive plant management.
Rangelands 32:43–47
49. Conant RT, Paustian K (2002) Potential soil carbon sequestration in overgrazed grassland
ecosystems. Glob Biogeochem Cycles 16(4):90-1–90-9
50. Boyte SP, Wylie BK, Guc Y, Major DJ (2020) Estimating abiotic thresholds for sagebrush
condition class in the Western United States. Rangel Ecol Manag 73:297–308
51. Brady NC, Weil RR (2016) Nature and properties of soils.15th edn. Pearson, Upper Saddle
River
52. Singh A (2021) Soil salinization management for sustainable development: a review. J Environ
Manag 277:111383
53. Schaetzl R, Thompson ML (2015) Soils: genesis and geomorphology.2nd edn. Cambridge
University Press, New York
54. Shahid SA, Zaman M, Heng L (2018) Guideline for salinity assessment, mitigation and
adaptation using nuclear and related techniques. Springer, Cham
55. Brown PL, Halvorson AD, Siddoway FH, Mayland HF, Miller MR (1982) Saline-seep diag-
nosis, control, and reclamation. U.S. Department of Agriculture Conservation Research Report
No. 30. US Government Printing Office, Washington DC
56. Halvorson AD, Reule CA (1980) Alfalfa for hydrologic control of saline seeps. Soil Sci Soc Am
J 44:370–374
57. Biederbeck VO, Bouman OT (1994) Water use by annual green manure legumes in dryland
cropping systems. Agron J 86:543–549
58. Obrycki JF, Karlen DL (2020) Forages for conservation and improved soil quality. In: Moore
KJ, Collins M, Nelson CJ, Redfearn DD (eds) Forages: the science of grassland agriculture,
II7th edn. Wiley, Chichester, pp 227–248
59. Havlin JL, Tisdale SL, Nelson WL, Beaton JD (2013) Soil fertility and fertilizers.8th edn.
Pearson-Prentice Hall, Upper Saddle River, NJ
60. Klaustermeier AW, Daigh ALM, Limb RF, Sedivec K (2017) Crystallization inhibitors and
their remediation potential on brine-contaminated soils. Vadose Zone J 16(4):1–10
61. Baird C, Cann M (2005) Environmental chemistry.3rd edn. W.H. Freeman and Company,
New York
62. Hazlett P, Emilson C, Lawrence G, Fernandez I, Ouimet R, Bailey S (2020) Reversal of forest
soil acidification in the Northeastern United States and Eastern Canada: site and soil factors
contributing to recovery. Soil Syst 4:54. https://doi.org/10.3390/soilsystems4030054
63. EPA (2020) Summary of the Clean Air Act. https://www.epa.gov/laws-regulations/summary-
clean-air-act. Accessed 15 Aug 2021
Agricultural Land Degradation in the United States of America 391

64. Barak P, Jobe BO, Krueger AR, Peterson LA, Laird DA (1997) Effects of long-term soil
acidification due to nitrogen fertilizer inputs in Wisconsin. Plant Soil 197:61–69
65. Jones C, Engel R, Olson-Rutz K (2019) Soil acidification in the semiarid regions of North
America’s Great Plains. Crops Soils. https://doi.org/10.2134/cs2019.52.0211
66. Reeves JL, Liebig MA (2016) Soil pH and exchangeable cation responses to tillage and
fertilizer in dryland cropping systems. Commun Soil Sci Plant Anal 47:2396–2404
67. Chien SH, Kallenbach RL, Gearhart MM (2010) Liming requirement for nitrogen fertilizer-
induced soil acidity: a new examination of AOAC guidelines. Better Crops 94(2):8–9
68. Wallace A (1994) Soil acidification from use of too much fertilizer. Commun Soil Sci Plant
Anal 25(1–2):87–92
69. Fanning DS, Rabenhorst MC, Fitzpatrick RW (2017) Historical developments in the under-
standing of acid sulfate soils. Geoderma 308:191–206
70. Soil Survey Staff (2014) Keys to soil taxonomy.12th edn. U.S. Department of Agriculture,
Natural Resources Conservation Service, Washington DC
71. Demas SY, Hall AM, Fanning DS, Rabenhorst MC, Dzantor EK (2004) Acid sulfate soils in
dredged materials from tidal Pocomoke Sound in Somerset County, MD, USA. Aust J Soil Res
42:537–545
72. SACPB (2003) Coastline–a strategy for implementing CPB policies on coastal acid sulfate soils
in South Australia. No 33. pp. 1–12
73. Fitzpatrick RW, Merry RH, Williams J, White I, Bowman G, Taylor G (1998) Acid sulfate soil
assessment: coastal, inland and minespoil conditions. National Land and Water Resources
Audit Methods Paper. 18 p
74. Mazmanian DA, Kraft ME (1999) The three epochs of the environmental movement. In:
Mazmanian DA, Kraft ME (eds) Towards sustainable communities: transition and transforma-
tions in environmental policy. The MIT Press, Cambridge, pp 3–41
75. Mortvedt JJ (1987) Cadmium levels in soils and plants from some long-term soil fertility
experiments in the United States of America. J Environ Qual 16:137–142
76. Brevik EC (2009) Soil, food security, and human health. In: Verheye W (ed) Soils, plant growth
and crop production. Encyclopedia of Life Support Systems (EOLSS), EOLSS Publishers,
Oxford. http://www.eolss.net. Accessed 15 Aug 2021
77. Burgess LC (2013) Organic pollutants in soil. In: Brevik EC, Burgess LC (eds) Soils and human
health. CRC Press, Boca Raton, pp 83–106
78. Brevik EC, Slaughter L, Singh BR, Steffan JJ, Collier D, Barnhart P, Pereira P (2020) Soil and
human health: current status and future needs. Air Soil Water Res 13:1–23

You might also like